Click through the PLOS taxonomy to find articles in your field.
For more information about PLOS Subject Areas, click here .
Loading metrics
Open Access
Peer-reviewed
Research Article
The effect of light quality on plant physiology, photosynthetic, and stress response in Arabidopsis thaliana leaves
Roles Conceptualization, Data curation, Formal analysis, Methodology, Visualization, Writing – original draft
* E-mail: [email protected] (ML); [email protected] (NY)
Affiliation Department of Bioresource Engineering, McGill University–Macdonald Campus, Sainte-Anne-de-Bellevue, Quebec, Canada
Roles Methodology
Affiliation Department of Plant Science, McGill University–Macdonald Campus, Sainte-Anne-de-Bellevue, Quebec, Canada
Roles Writing – review & editing
Roles Resources
Roles Conceptualization, Funding acquisition, Investigation, Project administration, Resources, Supervision, Writing – original draft
- Nafiseh Yavari,
- Rajiv Tripathi,
- Bo-Sen Wu,
- Sarah MacPherson,
- Jaswinder Singh,
- Mark Lefsrud
- Published: March 4, 2021
- https://doi.org/10.1371/journal.pone.0247380
- Reader Comments
The impacts of wavelengths in 500–600 nm on plant response and their underlying mechanisms remain elusive and required further investigation. Here, we investigated the effect of light quality on leaf area growth, biomass, pigments content, and net photosynthetic rate (Pn) across three Arabidopsis thaliana accessions, along with changes in transcription, photosynthates content, and antioxidative enzyme activity. Eleven-leaves plants were treated with BL; 450 nm, AL; 595 nm, RL; 650 nm, and FL; 400–700 nm as control. RL significantly increased leaf area growth, biomass, and promoted Pn. BL increased leaf area growth, carotenoid and anthocyanin content. AL significantly reduced leaf area growth and biomass, while Pn remained unaffected. Petiole elongation was further observed across accessions under AL. To explore the underlying mechanisms under AL, expression of key marker genes involved in light-responsive photosynthetic reaction, enzymatic activity of antioxidants, and content of photosynthates were monitored in Col-0 under AL, RL (as contrast), and FL (as control). AL induced transcription of GSH2 and PSBA , while downregulated NPQ1 and FNR2 . Photosynthates, including proteins and starches, showed lower content under AL. SOD and APX showed enhanced enzymatic activity under AL. These results provide insight into physiological and photosynthetic responses to light quality, in addition to identifying putative protective-mechanisms that may be induced to cope with lighting-stress in order to enhance plant stress tolerance.
Citation: Yavari N, Tripathi R, Wu B-S, MacPherson S, Singh J, Lefsrud M (2021) The effect of light quality on plant physiology, photosynthetic, and stress response in Arabidopsis thaliana leaves. PLoS ONE 16(3): e0247380. https://doi.org/10.1371/journal.pone.0247380
Editor: Keqiang Wu, National Taiwan University, TAIWAN
Received: August 31, 2020; Accepted: December 13, 2020; Published: March 4, 2021
Copyright: © 2021 Yavari et al. This is an open access article distributed under the terms of the Creative Commons Attribution License , which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Data Availability: All relevant data are within the manuscript and its Supporting Information files.
Funding: This work was supported by the “Natural Sciences and Engineering Research Council of Canada (NSERC)”. The specific grant number is RGPIN 355743-13, CRDPJ418919-11. It is “all” the funding and/or financial sources of support (whether external and/or internal to our organization) that were received during this study. And there was no additional external and/or internal funding received for this study. ML is the author, who received this award. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.
Competing interests: The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.
Introduction
Among various environmental factors, light is one of the most important variables affecting photosynthesis as well as plant growth and development [ 1 ]. Plants require light not only as an energy source but also as a clue to adjust their development to environmental conditions. During photosynthesis, absorbed energy is transferred to the photosynthetic apparatus, which is comprised of Photosystem I (PSI), Photosystem II (PSII), electron transport carriers (cytochrome b6f (cytb6f), plastoquinone (PQ), plastocyanin (PC)), and ATP synthase. The light-responsive photosynthetic process is driven by the released electrons through the water-splitting reaction on the PSII side, followed by NADP + reduction to NADPH, and proton flow into the lumen in order to generate ATP. Generated NADPH and ATP serve as an energy source for the carbon fixation process [ 2 ].
Both quality and quantity of incident light can have drastic impacts on photosynthetic activity and photosystem adaption to changing light quality [ 3 , 4 ]. Earlier studies on photosynthetic activity reported that photosynthesis is a wavelength-dependent response, in which amber light (AL; 595 nm) induces higher photosynthetic rates than blue light (BL; 450 nm) or red light (RL; 650 nm) [ 3 , 5 , 6 ]. These studies have become the foundation for our plant lighting research as light emitting diodes (LEDs) are proven to be an optimal and effective tool to study the effect of wavelength on plant physiological and biochemical responses [ 7 – 10 ]. Prior research has demonstrated that the wavelength range from 430–500 nm is effective at simulating pigmentation, metabolism of secondary metabolites, photosynthetic function, and development of chloroplasts [ 11 – 14 ]. The wavelength range of 640–670 nm was found effective in promoting photosynthetic activity, plant biomass and leaf area growth [ 3 , 15 ] while playing critically important roles in the development of photosynthetic apparatus, net photosynthetic rate (Pn) and primary metabolism [ 12 , 16 ]. Growing research on the wavelength range 500–600 nm have highlighted its important physiological and morphological impact on growth, chlorophyll content, and photosynthetic function [ 8 , 17 – 19 ]. However, conflicting results on the impact of AL were reported [ 3 , 20 ]. Although AL results in high photosynthetic activity, poor plant growth responses such as elongation and growth suppression have been reported [ 20 , 21 ], and this underlying mechanism remains unknown. In addition to this, AL is weakly absorbed by the photosynthetic pigments [ 22 ]. At the current state, further investigation of AL impact is required to better understand the photoactivity of the photosystems.
Recent studies reported that light quality and quantity can have drastic impacts on imbalanced excitation of either PSII or PSI, resulting in energy imbalance between photosystems and triggering stoichiometric adjustments of photosynthetic complexes [ 23 , 24 ]. This imbalance between the two photosystems can result in generation of harmful reactive intermediates, mainly reactive oxygen species (ROS) [ 25 , 26 ]. Generation of ROS can result in oxidative damage to the chloroplasts, leading to photosystem photo-inhibition that strongly limits plant growth [ 27 ]. To maintain steady state photosynthetic efficiency and prevent ROS accumulation, plants activate the buffering mechanisms, including cyclic photosynthetic electron flow (CEF) and non-photochemical quenching (NPQ) [ 28 , 29 ]. To scavenge ROS, plants further stimulate antioxidative mechanisms via enhanced activity of associated enzymes such as glutathione synthetase (GSH), ascorbate peroxidase (APX), and superoxide dismutase (SOD) [ 30 ]. These studies and their findings allow us to understand the impact of light within photosystems; however, the wavelength that can induce such stress responses and their physiological consequence on plants remain poorly studied.
Therefore, to better understand the effect of light quality on plant growth and photosynthetic performance, we studied three narrow-wavelength LEDs of blue light (BL; 450 nm), amber light (AL; 595 nm), and red light (RL; 650 nm), and compared them with fluorescent light (FL; 400–700) as the control. We chose light quality of BL and RL as leaf pigments have absorption peaks at these wavelengths [ 31 ]. AL was chosen due to the conflicting results between high photosynthetic activity and poor plant growth responses [ 3 , 5 ]. Furthermore, to assess whether light quality-induced changes in plant growth and photosynthesis are mediated by the genotype, we investigated the light quality response in three A . thaliana accessions Col-0, Est-1, and C24. These accessions show different geographical distributions and hence are adopted to different environments [ 32 – 34 ]. Congruently, they show a high degree of divergence in their photosynthetic response to the light environment [ 35 , 36 ]. Two experiments were designed to assess the impact of light quality on the plant. First, we investigated the physiological and photosynthetic response of A . thaliana to BL, AL, and RL lights compared it to FL by measuring leaf area growth, biomass content, Pn, and pigments content. Second, we tested whether changes in plant response to light quality is genotype specific by conducting the experiments across three A . thaliana accessions. Third, we investigated the potential induction of stress responses under AL by testing whether there are light quality-specific changes in the expression of marker genes involved in light-responsive photosynthetic process and enzymatic activity of antioxidants, as well as photosynthates content. Our findings expand the current understanding on physiological and photosynthetic responses of plants to light quality, in addition to identifying putative protective-mechanisms that may be induced to cope with lighting-stress in order to enhance plant stress tolerance.
Materials and methods
Plant materials and growth condition.
Seeds of A . thaliana accessions Col-0, Est-1, and C24 were obtained from the Arabidopsis Biological Resource Center (ABRC; Columbus, OH, US). Seeds were placed in rockwool cubes (Grodan A/S, DK-2640, Hedehusene, Denmark) and incubated at 4°C for 2 days. White broad-spectrum light (FL; 4200 K, F72T8CW, Osram Sylvania, MA, US) were used as light sources for seed germination. Seedlings were hydroponically grown under FL for 21 days with the environmental condition of 24 h photoperiod, 23°C, 50% relative humidity, and ambient CO 2 in a growth chamber (TC30, Conviron, Winnipeg, MB, Canada). Seed density was adjusted to limit treated plants from shadowing each other. FL was placed over the plant-growing surface area (49 cm × 95 cm) at a low photosynthetic photon flux density (PPFD) of 69 to 71 μmol·m -2 ·sec -1 . PPFD was measured at the conjunction of a grid (square area 3 cm 2 ) placed over the growing area. After 21 days, plants formed rosettes with nine (C24) and eleven (Col-0 and Est-1) leaves. To reach the same growth stage as Col-0 and Est-1 plants, C24 plants were allowed to grow for 23 days [ 37 ]. Fresh half-strength Hoagland nutrient solution [ 38 ] was provided every other day.
Light treatment
After day 21 (Col-0 and Est-1) or 23 (C24), plants were transferred to their respective light treatment for 5 days, each with the same environmental conditions: 24 h photoperiod, 23°C, 50% relative humidity, and ambient CO 2 . 21-day old plants were randomly divided into four experimental groups and received treatments using light emitting diodes (LED) (VanqLED, Shenzhen, China) of BL (peak wavelength: 450 nm), AL (peak wavelength: 595 nm), and RL (peak wavelength: 650 nm). The fourth group was treated with FL (400–700 nm), as the control. The light spectra and PPFD were monitored daily by using a PS-300 spectroradiometer (Apogee, Logan, UT, US). PPFD was maintained at 69 to 71 μmol·m -2 ·sec -1 throughout the whole plant growth period. Fresh half-strength Hoagland nutrient solution [ 38 ] was provided every other day. Biological replicates were grown at different time points under the same environmental settings.
Physical and biochemical analyses
Leaf area growth determination..
Three plants per biological replicate were randomly selected for each measurement. Leaves from the selected plants were collected for the determination after treatment (5 days). Digital images of leaves were taken with a window size of 640 x 480 pixels and a camera-object distance of approximately 80 cm. The digital images were next used to determine leaf area growth using Image J software with default settings (Bethesda, MD, US), as described previously [ 39 ].
Biomass content determination.
Three plants per biological replicate were randomly selected for each determination. Leaf samples from the selected plants were collected for the dry mass determination before (0 h) and after treatment (5 days). Leaves were dried at 80°C for 2 days until a constant mass was achieved (less than < 5% mass difference over a 2 h period).
Pigment content determination.
Five plants per biological replicate were randomly selected for each assay. Leaf samples from the selected plants were collected for the determination after treatment (5 days). Methods and equations described by [ 40 – 42 ] were used to estimate the content of chlorophyll (Chl a and Chl b), carotenoids, and anthocyanin, respectively. Briefly, chlorophylls and carotenoids were extracted with 5 ml of 80% acetone at 4°C overnight, before centrifugation at 13,000 g for 5 min. Total anthocyanins were determined by extracting with 5 ml 80% methanol containing 1% HCl solvent at 4°C overnight, before centrifugation at 13,000 g for 5 min. The absorbance of the extraction solution was determined for Chl a (664 nm), Chl b (647 nm), carotenoids (440 nm), and anthocyanins (530 nm and 657 nm) using a UV–VIS spectrophotometer (UV-180, Shimadzu, Japan).
Net photosynthetic rate determination
Net photosynthetic rate was monitored before (0 h) and after treatment (5 days) using the LI-6400XT Portable Photosynthesis System (LI-COR Biosciences, Lincoln, NE, US) equipped with a 6400–17 Whole Plant Arabidopsis Chamber (LI-COR Biosciences). To reduce potential measurement errors, three plants were grouped as a single sample for determinations [ 43 ]. To avoid mismatch between the light quality used by the LI-6400XT Portable Photosynthetic System, and the LED lights used for the treatments [ 44 ], measurements were taken inside the controlled-chamber, in which whole plants (still embedded in rockwool) were placed and illuminated with LEDs. As a precaution, parafilm was placed on top of the rockwool cube to maintain moisture within the root zone while measurements were recorded. The environmental conditions of the chamber were set as: 400 ppm CO 2 , 50% relative humidity, 23°C, and 400 μl min -1 flow rate. Each measurement was taken over 20 min, including 5 min in the dark and 10–15 min under a light treatment at 69–71 μmol·m -2 ·sec -1 . A stable Pn reading was reached 10 min after illumination. Leaf area growth was determined to normalize Pn per unit leaf area growth. Measurements for three replicates (three plants per replicate, three replicates per treatment) were performed.
Photosynthate content determination
Previous studies have reported that the diurnal cycle and developmental stage of plants, along with the stress response can affect the plant metabolism [ 45 , 46 ]. Thus, we performed a time course assessment of 0, 1, 3, 5, and 7 days to determine the content of leaf photosynthates (proteins, starches, and lipids). Five plants per biological replicate were randomly selected for each measurement. Leaf samples from selected plants were collected for the determination prior (0 h) and after light treatments (1, 3, 5, and 7 days). Samples were immediately frozen in liquid nitrogen and stored at -80 ∘ C, before they were used for determination. Protein : Total protein content was measured using the Pierce™ BCA Protein Assay Kit (Thermo Fisher Scientific, Rockford, IL, USA). As a standard, the absorbance of the bovine serum albumin was determined at UV/Vis: λ max 562 nm. Starch : A previously described method [ 47 ] was used to estimate total starch content. Lipids : Previously described methods [ 48 , 49 ] were used (with minor modifications) to estimate the total lipid content. Briefly, each sample was homogenized with (CHCl 3 /MeOH, 70:30 v/v), before centrifuged at 1000 rpm for 5 min. The collected supernatant was incubated for 30 min at 70°C in a boiling water bath. Next, (H 2 SO 4 : 1 ml) was added and heated for 20 min. Following 2 min cooling on ice, (H 3 PO 4 : 1.5 ml) was added and incubated for 10 min until a pink color developed.
Antioxidative enzyme activity estimation
Five plants per biological replicate were randomly selected for each measurement. Leaf samples from selected plants were collected for the determination after treatment (5 days). Samples were immediately frozen in liquid nitrogen and stored at -80 ∘ C, before they were used for determination. Methods described by [ 50 , 51 ] were used to monitor the activity of SOD and APX antioxidative enzymes, respectively. Enzymatic activity was measured for 5 min at room temperature. The protein content in the supernatant was determined by the Pierce™ BCA Protein Assay Kit. The activity of SOD and APX was expressed as unit min −1 mg −1 protein.
Gene expression analysis
Cdna synthesis..
Changes in transcription of the interested genes were analyzed in A . thaliana Col-0 treated for 24 h under AL, RL, and FL. Leaf samples from selected plants were collected for the determination prior to treatment (0 h) and after treatment (2 h, 4 h, and 24 h). Samples were immediately frozen in liquid nitrogen and stored at -80 ∘ C, before they were used for determination. Four biological replicates were examined. For each biological replicate, five A . thaliana plants were selected, and their leaves were pooled together to represent a biological replicate. Plants in each biological replicate were grown independently, and at different times. Total RNA was extracted from (100 mg) leaves using the Sigma Spectrum Plant Total RNA Kit (STRN50; Sigma, Seelze, Germany) according to the manufacturer’s protocol. A total of (2 μg) RNA per sample was treated with amplification grade DNase I (Invitrogen, Carlsbad, CA, USA) to remove any traces of genomic DNA contamination. RNA concentrations were measured before and after DNase I digestion with a NanoDrop ND-1000 UV-Vis spectrophotometer (NanoDrop Technologies, Wilmington, Delware, USA). The cDNA was synthesized using AffinityScript QPCR cDNA Synthesis Kit (Agilent, Tech., Santa Clara, USA).
Primer design.
Primers for genes of interest ( S1 Table ) were designed using IDT software ( https://www.idtdna.com/calc/analyzer ) with the following criteria: Tm of 58–60°C and PCR amplicon lengths of 70 to 120 bp, yielding primer sequences 20 to 25 nucleotides in length with G-C contents of 40% to 50%. Specificity of the resulting primer pair sequences was examined using Arabidopsis transcript database with TAIR BLAST ( http://www.arabidopsis.org/Blast/ ). Specificity of the primer amplicons was further confirmed by melting-curve analysis (30 amplification cycles by PCR and subsequent gel-electrophoretic analysis). Primer amplicons were resolved on (agarose gels, 2% w/v) run at 110 V in Tris-borate/EDTA buffer, along with a 1Kb + DNA-standard ladder (Invitrogen, Carlsbad, CA, USA).
Quantitative real time-PCR (qRT-PCR) analysis.
Real-time qRT-PCR was performed with a MX3000P qPCR System (Agilent, Tech., Santa Clara, CA, USA) using three biological and two technical replicates, as described previously [ 52 ]. Relative expression was conducted following the manufacturer’s recommendations with two reference genes gamma tonoplast intrinsic protein 2 ( TIP2 ; AT3g26520) and actin 2 ( ACT2 ; AT3g18780) and the Brilliant III SYBR Green QPCR master mix (Agilent, Tech., Santa Clara, CA, USA). Amplification was performed in a (20 μL) reaction mixture containing (160 nmol) for each primer, 1x Brilliant III SYBR Green QPCR master mix, (15μM) ROX reference dye, and (0.3 μL) of cDNA template. Amplification conditions were 95°C for 10 min (hot start), followed by 40 cycles at 94°C for 30 s, 60°C for 30 s, and 72°C for 30 s. Fluorescence readings were taken at 72°C, at the end of the elongation cycle.
Data analysis.
Ct values were calculated with CFX-Manager and MX-3000P software. Relative expression changes (delta-delta Ct) were calculated according to [ 53 ] using A . thaliana TIP2 (AT3g26520) and ACT2 (AT3g18780) as reference genes. To avoid multiple testing, the p-values were only considered for 0 h with 24 h (a total of 12 genes and two light conditions). A gene was considered differentially expressed if p < 0.05 and the fold change pattern at 24 h was consistent with those observed at 2 and 4 h.
Statistical analysis
Differences between light treatments were tested using the two-tailed Student’s t-test. A two-way ANOVA was used to assess the effects of accession and different light treatments on leaf area growth, biomass content, Pn value, and pigments content. We observed similar patterns using the non-parametric tests of Wilcoxon-Mann-Whitney and Kruskal-Wallis tests (data not shown).
Effect of light quality and natural genotype variation on leaf area growth, biomass content, net photosynthetic rate, and pigment content in A . thaliana
To assess the effect of light quality, 21-days-old plants (11 leave plants) of three A . thaliana accessions Col-0, Est-1, and C24 were randomly divided into groups and treated under narrow-spectrum light (BL, AL, and RL), along with FL as control (baseline), for 5 days at approximately 70 μmol m -2 sec -1 ( Fig 1A and 1B ). Summary of light quantity compositions emitted from FL and LEDs light sources are shown in Table 1 . After 5 days of narrow-spectrum light treatments, leaf area growth, leaf biomass (dry mass), net photosynthetic rate, and pigment contents were measured across three A . thaliana accessions and compared with the baseline FL treatment ( Fig 1C–1E and Table 2 ).
- PPT PowerPoint slide
- PNG larger image
- TIFF original image
(A) Light emission spectra of LED light sources and FL. (B) Eleven-leaves stage A . thaliana accessions Col-0, Est-1, and C24 were grown hydroponically and treated for 5 days under narrow-spectrum BL, AL, and RL lights, as well as FL as control. (C) Leaf area growth. (D) Leaf biomass (dry mass). (E) Net photosynthetic rate (Pn) measured at 69–71 μmol m -2 sec -1 . Data are expressed as mean values ± standard deviation (n = 3). Statistical analysis was performed against FL using a two-tailed Student’s t-test (n.s.: not statistically significant; *: P < 0.05; **: P < 0.01).
https://doi.org/10.1371/journal.pone.0247380.g001
https://doi.org/10.1371/journal.pone.0247380.t001
https://doi.org/10.1371/journal.pone.0247380.t002
Under RL, the leaf area growth was significantly increased across accessions ( P < 0.05; Fig 1C ). Under BL, leaf area growth was significantly increased in C24 and Col-0 ( P < 0.05; Fig 2C ), but the increase in the leaf area growth was not significant in Est-1. Under AL, leaf area growth showed a severe reduction in Col-0 and C24 ( P < 0.05; Fig 1C ), while Al showed no change in Est-1. Petioles were noticeably elongated under AL ( Fig 1B ).
https://doi.org/10.1371/journal.pone.0247380.g002
The leaf biomass significantly increased under RL across the three accessions ( P < 0.05; Fig 1D ). Under BL, the leaf biomass was significantly decreased in Est-1 and C24 but increased in Col-0 ( P < 0.01; Fig 1D ). Under AL, the leaf biomass was significantly lower in Col-0 and C24 ( P < 0.01), while it showed no change in Est-1 ( Fig 1D ).
As for the net photosynthetic rates (Pn), it significantly increased under RL across the accessions ( P < 0.05; Fig 1E ). In contrast, there was no significant difference in Pn under AL ( Fig 1E ). Under BL, Pn significantly increased in Col-0 and Est-1 ( P < 0.05; Fig 1E ) but remained unchanged in C24.
There was no significant difference in contents of chlorophyll a (Chl a) and chlorophyll b (Chl b) in Col-0 and C24 under the light quality of BL, AL, and RL ( Table 2 ). In contrast, Chl a content significantly increased in Est-1 under RL ( P < 0.05; Table 2 ). Across accessions, Chl a: b content significantly increased, remained unchanged, and decreased under RL, BL, and AL, respectively ( Table 2 ). Moreover, there was no significant difference in carotenoid and anthocyanin contents across the accessions under AL and RL. However, BL significantly stimulated carotenoids content in Est-1 and Col-0 ( P < 0.05; Table 2 ). Additionally, anthocyanins content significantly increased under BL in Est-1 and C24 ( P < 0.01; Table 2 ).
The two-way ANOVA analysis indicated significant effects of the light treatments for the determined parameters, except Chl b. Also, the interaction between light treatments and genotype was significant for leaf area growth and leaf biomass ( P < 0.01; Table 3 ).
https://doi.org/10.1371/journal.pone.0247380.t003
Changes in transcription of photosynthetic marker genes, content of photosynthates, and activity of antioxidant in A . thaliana Col-0 under AL and RL
The severe reduction in leaf area growth and biomass, along with unchanged levels of Pn in Col-0 and C24 under AL suggested that amber light has mismatched effects on photosynthetic activity and photomorphology. Further to this, although chlorophyll contents under AL were 10–20% lower than the FL, both light treatments triggered similar photosynthetic activity, which implies that amber light has unidentified mechanisms in the photosynthetic process. To identify the mechanisms that amber light triggers within plants, we next explored transcriptional changes in marker genes associated with the photosynthetic light reaction and photo-protective mechanisms, photosynthates content and antioxidant enzymatic activity in Col-0 under AL ( Fig 2 ). Among three accession, Accession Col-0 was chosen for the transcription analysis, as it is the most common A . thaliana accession in conducting biological analysis. In addition to AL and FL (as control), changes were investigated under RL, as RL-treated plants showed opposing changes in leaf physiological phenotypes compared to AL.
Gene expression analysis indicated a significant increase in transcription level of ATP synthase gamma chain 1 ( ATPC1 ;member of ATP synthase complex) and proton gradient regulation Like 1 ( PGRL1B ;member of CET complex), after 24 h treatment under AL ( P < 0.05; Fig 3B ). ATPC1 transcription significantly increased after 24 h treatment under RL ( P < 0.05; Fig 3B ). No significant difference, after 24 h treatment, was observed in the transcription level of the selected marker genes associated with linear photosynthetic electron transfer (i.e., ferredoxin-2 ( Fd2) , plastocyanin (PETE1) , and cytochrome b6f complex ( PETC ) under both AL and RL ( Fig 3B ). After the 24 h treatment, transcription of ferredoxin-NADP+-oxidoreductase (FNR2) was significantly decreased under AL ( P < 0.05; Fig 3B ), while it remained unchanged under RL ( Fig 3B ). The transcription level of ribulose bisphosphate carboxylase small chain (RBCS1A) was significantly reduced at 2 h and 4 h treatment under both AL and RL ( P < 0.05; Fig 3B ). The RBCS1A transcription level significantly was downregulated under AL ( P < 0.05; Fig 3B ). However, RBCS1A transcription level recovered after the 24 h treatment under RL ( Fig 3B ).
(A) Genes of interest are highlighted in green. (B) Transcription of genes implicated in the light-responsive photosynthetic process that is located within the thylakoid membrane. A time course assessment prior to treatment (0 h), and after treatment (2, 4, and 24 h) of AL and RL was performed, compared to FL. Four biological replicates were examined. For each biological replicate, five A . thaliana plants were selected, and their leaves were pooled together to represent a biological replicate. Plants in each biological replicate were grown independently, and at different time.
https://doi.org/10.1371/journal.pone.0247380.g003
All data were normalized to the housekeeping genes; gamma tonoplast intrinsic protein 2 ( TIP2 ; AT3g26520) and actin 2 ( ACT2 ; AT3g18780). Red borders represent significant changes in expression ( P < 0.05). Studied genes include: ATP synthase gamma chain 1, ATPC1 (AT4g04640); fatty acid desaturase 6, FAD6 (AT4g30950); ferredoxin-2, Fd2 (AT1g60950); ferredoxin-NADP+-oxidoreductase, FNR2 (AT1g20020); (Fdx)-thioredoxin (Trx)-reductase, FTRB (AT2g04700); glutathione synthetase, GSH2 (AT5g27380); PSII nonphotochemical quenching, NPQ1 (AT1g08550); cytochrome b6f complex (Cyt b6f), PETC (AT4g03280); plastocyanin, PETE1 (AT1g76100); proton gradient regulation Like 1, PGRL1B (AT4g11960); photosystem II protein D1, PSBA (ATCG00020) and ribulose bisphosphate carboxylase small chain, RBCS1A (AT1g67090).
To confirm changes in the ATP synthase and CET complex under AL, we leveraged available proteomics data where eleven-leaves plants of A . thaliana Col-0 were grown under AL and RL for 5 days. Consistent with the observed transcriptomic data, a significant increase in the level of protein abundance was observed for both CET complex ( P < 1.3 x 10 −12 ; S1A Fig ) and ATP synthase ( P < 2 x 10 −4 ; S1B Fig ) under AL compared to RL.
Regulation patterns of PSBA , NPQ1 , GSH2 and FAD6 transcripts in A . thaliana Col-0 under AL and RL
The transcription level of photosystem II protein D1 ( PSBA) was significantly upregulated at 4 h and 24 h treatment under RL ( P < 0.05; Fig 3B ). Under AL, the transcription level of PBSA showed a similar increase after the 4 h treatment ( P < 0.05; Fig 3B ); However, its transcription level was reduced to a comparable level with FL after the 24 h treatment under AL. After the 24 h treatment, the transcription level of PSII nonphotochemical quenching (NPQ1) was significantly downregulated under AL ( P < 0.05; Fig 3B ), while it remained steady under RL. Between the 2 h and 4 h treatment, the transcription level of GSH2 gradually increased under both AL and RL ( P < 0.05; Fig 3B ) but reduced to a comparable level with FL after the 24 h treatment under RL. No significant difference was observed in the transcription level of fatty acid desaturase 6 ( FAD6) after the 24 h treatment under either AL or RL ( Fig 3B ).
Photosynthates content in A . thaliana Col-0 under AL and RL
Photosynthates accumulation was probed in Col-0 treated under AL, RL, and FL. Total lipid, protein and starch were measured at days 0, 1, 3, 5, and 7 ( Fig 4 ). The lipid content gradually increased under AL and RL ( P < 0.05; Fig 4A ). The content level of proteins and starches increased under RL but decreased under AL ( P < 0.05; Fig 4B and 4C ).
(A) Lipid; (B) Protein; (C) Starch. Data are expressed as mean values ± standard deviation (n = 5). Statistical analysis was performed against FL using a two-tailed Student’s t-test.
https://doi.org/10.1371/journal.pone.0247380.g004
Antioxidative enzyme activity in A . thaliana Col-0 under AL and RL
We examined the antioxidative activity of superoxide dismutase (SOD) and ascorbate peroxidase (APX) enzymes in Col-0 treated under AL, RL, and FL ( Fig 5 ). After the 24 h treatments, activity of both antioxidants was significantly increased under AL ( P < 0.05; Fig 5 ), while no significant changes were observed for either of these enzymes when plants were treated under RL.
(A) Superoxide dismutase (SOD) activity. One unit of SOD activity was defined as the amount of enzyme required to result in a 50% inhibition of the rate of reduction at 550 nm in 1 min. (B) Ascorbate peroxidase (APX) activity. One unit of APX activity was defined as the amount of enzyme required to oxidize 1 μmol of ascorbate at 290 nm in 1 min. Enzymatic activity was measured for 5 min at room temperature and data are expressed as mean values ± standard deviation (n = 5). Statistical analysis was performed against FL using a two-tailed Student’s t-test (n.s., not statistically significant; *, P < 0.05; **, P < 0.01).
https://doi.org/10.1371/journal.pone.0247380.g005
In this work, we investigated the impact of light quality BL, AL, and RL on leaf growth and photosynthetic response across three A . thaliana accessions Col-0, Est-1, and C24. The analyses clearly demonstrate the significant impact of light quality on leaf area growth, biomass content, and pigments accumulation (chlorophylls, carotenoid, and anthocyanin). The results indicate that light quality significantly influences Pn across accessions, consistent with the reported results that leaf photosynthetic reaction is wavelength-dependent in higher plants [ 54 ].
Importance of geographic habitats on light quality response of leaf growth and biomass
The selected accessions Col-0, Est-1, and C24 have different geographic habitats; C24 originated from a part of Europe (Portugal), Est-1 from Northern Asia (Russia), and Col-0 from United States (Columbia). Therefore, we took into account differences in geographical range for these accessions resulted in a high degree of divergence in photosynthetic characteristics to light [ 35 ]. The most extreme responder in leaf area growth and leaf biomass analyses was Est-1 from Russia. It is worth pointing out that the two Col-0 and C24 accessions highlighted here as weak responders, they elongated very quickly under AL and thus may not be true candidates for weak responders. Previous studies have found negative correlations between hypocotyl height and latitude of accession origin in European Arabidopsis accessions [ 55 , 56 ], suggesting that this natural variation in light sensitivity could be a result of adaptation to the north-south gradient in ambient light intensity.
The results of the study described here emphasize the strength of explicitly incorporating LxG interactions into the leaf area growth and leaf biomass content across the accessions. Importantly, as further elaborated below, the genotype-specific responses in leaf area growth and biomass content were observed exclusively under AL and BL, while the three accessions exhibited similar patterns of changes under RL. Our findings are consistent with previous reports on different accessions and light quality treatments, and underscores the importance of considering the natural habitat effect in characterizing the impact of light quality on leaves [ 57 ].
Leaf development varied between accessions such that the overall dynamic of growth and biomass were different. For example, we took efforts in synchronizing leaf growth stage in the accessions, resulting in the C24 plants being grow for 23 days to reach the same leaf stages of the plant. Some of the observed variation in leaf growth response could be simply a manifestation of the different time-course between accessions. These differences between accession can be significant and have the potential to enhance our understanding of the ecological role of specific adaptations.
Findings on BL supports its role on activation of protective pigments
BL induced higher leaf area growth across three accessions. However, its impact on biomass production is accession-dependent, and may be caused by accessory pigment accumulation (anthocyanins). Under BL, only Col-0 showed an increase in biomass, as opposed to Est-1 and C24, which showed a decrease in biomass. It was observed that BL induced a significantly higher concentration of anthocyanins in Est-1 and C24 than Col-0. These results imply that the impact of wavelength on accessory pigment accumulation is accession-dependent, and that this difference in accessory pigment accumulation consequently leads to differences in biomass production across accessions. Anthocyanin is a photo-protective pigment, which protects plant and its chloroplast membrane by absorbing blue light and against photo-oxidation [ 58 , 59 ]. Higher concentration of anthocyanin accumulating in a plant results in lower BL interception, which consequently lead to lower biomass production over the long term. Further to this, in this study, we found the ratio of Chl a:b is similar under BL and FL across accessions. This consistency in Chl a:b, suggests a lack of photosystems reconfiguration under BL [ 60 , 61 ]. Our results confirm the role of BL in stimulating anthocyanin content in plants and protecting them from light stress [ 62 ]. Plants activate photo-protective mechanisms under BL to cope with a potential induced-light stress, resulting in an increased accumulation of photo-protective pigments [ 58 , 59 ]. Notably, we found different patterns in content of anthocyanin accumulation in the accessions. Results showed that anthocyanin accumulation can be triggered at low BL (~70 μmol m -2 sec -1 ), which suggests that this protective mechanism against BL can vary based on the accession (i.e. natural adaptations) and can be triggered under low light. Further investigation on these two accessions on BL with a wide range of BL intensity is required. Our results thus encourage future studies analyzing this trait using BL with a wide range of BL intensity to further advance our understanding of the underlying mechanisms.
Plants showed high antioxidative and photo-protective under AL
AL had no impact on the photosynthetic activity across the three accessions compared to FL; yet it induced the poorest morphological traits. Col-0 and C24 showed a severe reduction in leaf area growth and biomass, while Est-1 was unaffected. These two accessions (Col-0 and C24) showed a clear elongation of petioles under AL, which suggests that leaf resources are redirected from leaves to petioles as insufficient lighting conditions under AL were performed in this study [ 63 ]. However, the results on transcriptional changes and photosynthates content showed the opposite responses to the morphological traits.
The photosynthates, including proteins and starches, showed lower content in leaves of plants treated under AL. A downregulation of RBCS1A (small subunit of Rubisco) transcription was also observed in the leaves treated under AL. A lower accumulation of proteins was previously observed under AL [ 64 ], suggesting a positive contribution of downregulated Rubisco genes, as it is the main protein in leaves. A lower content of carbohydrates under stress conditions has been observed before in A . thaliana [ 65 ]. Future work is needed to explore if a reduced conversion of light energy into chemical energy has occurred in the photosynthesis process under AL.
High capacity for lipids accumulation was observed for plants treated under AL. Lipid accumulation had been previously linked to oxidative stress [ 66 ]. suggesting an increase in lipophilic antioxidants content such as tocopherols, which play an important role in the scavenging of singlet oxygen [ 67 ]. Moreover, we found a significant increase in both expression and enzymatic activity of antioxidants under AL. Plants stimulate antioxidative mechanisms to protect the photosynthetic apparatus from incurring damage via ROS detoxification [ 30 , 68 ]. Our results on photosynthates thus suggest that plants tried to cope with a potential ROS stress condition under AL.
A significant upregulation in glutathione biosynthesis, transcription level of PGRL1B , ATPC1 , and marker genes associated with ATP synthase and CET complex was observed. In agreement with this result, a significant increase in the expression of ATPC1 at the protein level was recently reported in A . thaliana Col-0 treated with 595 nm light [ 69 ]. CET plays an important role to protect plants under high light and other stress environments [ 70 ]. During CET, electrons are cycled around PSI and protons are translocated to generate a proton gradient across the thylakoid membranes [ 71 ]. In addition to contributing ATP synthesis, another function of a generated proton gradient is to dissipate excess energy as heat from the PSII antennae [ 72 ]. Further to this, an upregulation of CET and ATP synthase suggests of an accelerated rate of PSII repair through elevated ATP synthesis [ 73 , 74 ]. As such, the results on photosynthates and at the transcription level under AL both suggest that AL, even at low light, induces protective mechanisms of photosystems related to light stress, which consequently triggers low protein and starch accumulation and result in poor morphological traits.
One possible hypothesis for the conflicting AL responses can be explained by the detour effect [ 75 , 76 ], where a major part of AL transmitted into the leaf is reflected within leaf tissues and re-absorbed by unsaturated chlorophylls multiple times, which leads to an observed light stress response. Due to the nature of the high absorbing efficiency of the chloroplast, nearly 90% of BL and RL are absorbed at the leaf surface and their detour effect is small [ 76 , 77 ]. While for the wavelength within 500–600 nm [i.e. green light (GL) and AL] that are less absorbed by chloroplast, its light path can increase by several folds and this results in its increased/overexpressed photosynthetic activity through light absorption by unsaturated chloroplast. Although there is no study reporting the underlying mechanisms triggered by AL, several studies have observed the impact of supplemented GL and AL on photosynthetic activity and plant productivity in horticultural plants, which reinforces our hypothesis on the increased photosynthetic activity under AL. Further to this, the aggressive suppression responses on morphological traits in A . thaliana under AL, opposed to the positive impact on plant development, is expected as A . thaliana is a low light plant. They are more sensitive to the change in light properties. Overall, our results suggest AL as a potential light source in activating the potential of increased plant productivity efficiently, but it requires high control on its intensity. This study clarifies why AL alone induces overexpressed high photosynthetic activity yet results in poor plant development.
RL modulated plant adaptation and energy assimilation
The leaf area growth was significantly increased under RL across all accessions, which in turn enabled a greater light interception by the leaves [ 78 ]. This agrees with the increased Pn that was observed across accessions. These observations along with a significant increase of leaf biomass suggests proper plant adaptation under RL across accessions. We found a significant increase in the Chl a: b under RL across accession. Chl a is mainly concentrated around PSI and PSII, whereas Chl b is most abundant in light-harvesting complexes [ 79 ]. An increase in Chl a: b can increase the likelihood of an efficient electron transfer system within the chloroplast membrane [ 80 ]. This, in turn, could positively influence the photosynthetic performance in plants under RL. Considering that timely synthesis of D1 protein is key to maintain the PSII function and consequently, photosynthetic performance in leaves [ 25 ]. An increasing trend of PSBA expression was observed in plants under RL. The PSBA gene is critical for the de novo synthesis of the D1 protein during PSII repairs [ 81 , 82 ]. Therefore, upregulated transcription of PSBA gene could play an important role in accelerating the process of D1 protein turnover under RL. Plants showed that leaf photosynthates (starches, lipids, proteins) increased under RL. Overall, our results present RL as an efficient light source in helping the leaf energy assimilation process, resulting in an increased leaf growth, photosynthetic performance, and photosynthates content in plants.
Supporting information
https://doi.org/10.1371/journal.pone.0247380.s001
S1 Table. List of primers sequences used in qPCR experiments.
https://doi.org/10.1371/journal.pone.0247380.s002
S1 Fig. Proteins involved in ATP synthase and CET complex of A . thaliana Col-0 are upregulated under AL (595 nm) compared to RL (650 nm).
In this experiment, eleven-leaves plants were grown under AL and RL for 5 days (three biological replicates per light condition). A) The expression pattern of protein members involved in Cyclic electron transfer (CET) complex. B) The expression pattern of protein members involved in ATP synthase complex. Expression levels for each protein is normalized to have mean of zero and standard deviation of one. Yellow or blue color indicates upregulation or downregulation, respectively.
https://doi.org/10.1371/journal.pone.0247380.s003
- View Article
- PubMed/NCBI
- Google Scholar
- 31. <oj/>E. W. Schwieterman, "Surface and temporal biosignatures," arXiv preprint arXiv:1803.05065, 2018.
- 38. Hoagland D. R. and Arnon D. I., The water-culture method for growing plants without soil . Berkely,: The College of Agriculture, 1950, p. 32 p.
- 54. Hoover W. H., The dependence of carbon dioxide assimilation in a higher plant on wave length of radiation . City of Washington,: The Smithsonian institution, 1937, p. 1 p.
Information
- Author Services
Initiatives
You are accessing a machine-readable page. In order to be human-readable, please install an RSS reader.
All articles published by MDPI are made immediately available worldwide under an open access license. No special permission is required to reuse all or part of the article published by MDPI, including figures and tables. For articles published under an open access Creative Common CC BY license, any part of the article may be reused without permission provided that the original article is clearly cited. For more information, please refer to https://www.mdpi.com/openaccess .
Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications.
Feature papers are submitted upon individual invitation or recommendation by the scientific editors and must receive positive feedback from the reviewers.
Editor’s Choice articles are based on recommendations by the scientific editors of MDPI journals from around the world. Editors select a small number of articles recently published in the journal that they believe will be particularly interesting to readers, or important in the respective research area. The aim is to provide a snapshot of some of the most exciting work published in the various research areas of the journal.
Original Submission Date Received: .
- Active Journals
- Find a Journal
- Journal Proposal
- Proceedings Series
- For Authors
- For Reviewers
- For Editors
- For Librarians
- For Publishers
- For Societies
- For Conference Organizers
- Open Access Policy
- Institutional Open Access Program
- Special Issues Guidelines
- Editorial Process
- Research and Publication Ethics
- Article Processing Charges
- Testimonials
- Preprints.org
- SciProfiles
- Encyclopedia
Article Menu
- Subscribe SciFeed
- Recommended Articles
- PubMed/Medline
- Google Scholar
- on Google Scholar
- Table of Contents
Find support for a specific problem in the support section of our website.
Please let us know what you think of our products and services.
Visit our dedicated information section to learn more about MDPI.
JSmol Viewer
How light reactions of photosynthesis in c4 plants are optimized and protected under high light conditions.
1. Diversity of C4 Photosynthesis
2. ways of light energy utilization: balanced distribution between photosystems and emission of excess energy as a heat, 2.1. elevation of cyclic electron transport components.
Click here to enlarge figure
2.2. Function of PTOX Protein and Chlororespiration
2.3. changes in the amount of thylakoid complexes and rearrangement of super- and megacomplexes, 2.4. photoinhibition and role of d1 protein phosphorylation, 2.5. state transitions and phosphorylation of lhcii, 2.6. xanthophyll cycle and heat dissipation, author contributions, institutional review board statement, informed consent statement, data availability statement, acknowledgments, conflicts of interest, abbreviations.
- Yamori, W.; Hikosaka, K.; Way, D.A. Temperature response of photosynthesis in C 3 , C 4 , and CAM plants: Temperature acclimation and temperature adaptation. Photosynth. Res. 2014 , 119 , 101–117. [ Google Scholar ] [ CrossRef ] [ PubMed ]
- Hatch, M.D.; Slack, C.R. Photosynthesis by sugar-cane leaves. A new carboxylation reaction and the pathway of sugar formation. Biochem. J. 1966 , 101 , 103–111. [ Google Scholar ] [ CrossRef ] [ PubMed ]
- Slack, C.R.; Hatch, M.D. Comparative studies on the activity of carboxylases and other enzymes in relation to the new pathway of photosynthetic carbon dioxide fixation in tropical grasses. Biochem. J. 1967 , 103 , 660. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
- Osmond, C.B. Beta-carboxylation during photosynthesis in Atriplex. Biochim. Biophys. Acta 1967 , 141 , 197–199. [ Google Scholar ] [ CrossRef ]
- Hatch, M.D. C4 photosynthesis: A unique elend of modified biochemistry, anatomy and ultrastructure. Biochim. Biophys. Acta Rev. Bioenerg. 1987 , 895 , 81–106. [ Google Scholar ] [ CrossRef ]
- Gutierrez, M.; Gracen, V.E.; Edwards, G.E. Biochemical and cytological relationships in C 4 plants. Planta 1974 , 119 , 279–300. [ Google Scholar ] [ CrossRef ]
- Hatch, M.; Kagawa, T.; Craig, S. Subdivision of C 4 -Pathway Species Based on Differing C 4 Acid Decarboxylating Systems and Ultrastructural Features. Funct. Plant Biol. 1975 , 2 , 111–128. [ Google Scholar ] [ CrossRef ]
- Furbank, R.T. Evolution of the C4 photosynthetic mechanism: Are there really three C4 acid decarboxylation types? J. Exp. Bot. 2011 , 62 , 3103–3108. [ Google Scholar ] [ CrossRef ] [ Green Version ]
- Edwards, G.E.; Voznesenskaya, E.V. Chapter 4 C 4 Photosynthesis: Kranz Forms and Single-Cell C 4 in Terrestrial Plants. In C 4 Photosynthesis and Related CO 2 Concentrating Mechanisms ; Springer: Dordrecht, The Netherlands, 2011; pp. 29–61. [ Google Scholar ] [ CrossRef ]
- Walker, B.J.; Kramer, D.M.; Fisher, N.; Fu, X. Flexibility in the Energy Balancing Network of Photosynthesis Enables Safe Operation under Changing Environmental Conditions. Plants 2020 , 9 , 301. [ Google Scholar ] [ CrossRef ] [ Green Version ]
- Hatch, M.D. C 4 Photosynthesis: An Unlikely Process Full of Surprises. Plant Cell Physiol. 1992 , 33 , 333–342. [ Google Scholar ] [ CrossRef ]
- Hatch, M.D.; Osmond, C.B. 4. Compartmentation and Transport in C 4 Photosynthesis. In Encyclopedia of Plant Physiology, New Series Vol 3, Transport in Plants III ; Stocking, C.R., Heber, U., Eds.; Springer: Berlin/Heidelberg, Germany; New York, NY, USA, 1976; pp. 144–184. [ Google Scholar ]
- Yoshimura, Y.; Kubota, F.; Ueno, O. Structural and biochemical bases of photorespiration in C 4 plants: Quantification of organelles and glycine decarboxylase. Planta 2004 , 220 , 307–317. [ Google Scholar ] [ CrossRef ] [ PubMed ]
- Fisher, D.D.; Schenk, H.J.; Thorsch, J.A.; Ferren, W.R. Leaf anatomy and subgeneric affiliations of C 3 and C 4 species of Suaeda (Chenopodiaceae) in North America. Am. J. Bot. 1997 , 84 , 1198–1210. [ Google Scholar ] [ CrossRef ] [ PubMed ]
- Voznesenskaya, E.V.; Franceschi, V.R.; Pyankov, V.I.; Edwards, G.E. Anatomy, chloroplast structure and compartmentation of enzymes relative to photosynthetic mechanisms in leaves and cotyledons of species in the tribe Salsoleae (Chenopodiaceae). J. Exp. Bot. 1999 , 50 , 1779–1795. [ Google Scholar ] [ CrossRef ]
- Voznesenskaya, E.V.; Franceschi, V.R.; Chuong, S.D.X.; Edwards, G.E. Functional characterization of phosphoenolpyruvate carboxykinase-type C 4 leaf anatomy: Immuno-, cytochemical and ultrastructural analyses. Ann. Bot. 2006 , 98 , 77–91. [ Google Scholar ] [ CrossRef ] [ Green Version ]
- Furbank, R.; Hatch, M. Mechanism of C 4 Photosynthesis: The Size and Composition of the Inorganic Carbon Pool in Bundle Sheath Cells. Plant Physiol. 1987 , 85 , 958–95864. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
- Burnell, J.N.; Hatch, M.D. Low Bundle Sheath Carbonic Anhydrase Is Apparently Essential for Effective C 4 Pathway Operation. Plant Physiol. 1988 , 86 , 1252–1256. [ Google Scholar ] [ CrossRef ] [ Green Version ]
- Sage, R.F.; Christin, P.-A.; Edwards, E.J. The C 4 plant lineages of planet Earth. J. Exp. Bot. 2011 , 62 , 3155–3169. [ Google Scholar ] [ CrossRef ]
- Sage, R.F.; Stata, M. Photosynthetic diversity meets biodiversity: The C 4 plant example. J. Plant Physiol. 2015 , 172 , 104–119. [ Google Scholar ] [ CrossRef ]
- Brown, R.H. A Difference in N Use Efficiency in C3 and C4 Plants and its Implications in Adaptation and Evolution. Crop Sci. 1978 , 18 , 93–98. [ Google Scholar ] [ CrossRef ]
- Sage, R.F.; Pearcy, R.W.; Seemann, J.R. The Nitrogen Use Efficiency of C3 and C4 Plants: III. Leaf Nitrogen Effects on the Activity of Carboxylating Enzymes in Chenopodium album (L.) and Amaranthus retroflexus (L.). Plant Physiol. 1987 , 85 , 355. [ Google Scholar ] [ CrossRef ] [ Green Version ]
- Jordan, D.B.; Ogren, W.L. The CO 2 /O 2 specificity of ribulose 1,5-bisphosphate carboxylase/oxygenase. Planta 1984 , 161 , 308–313. [ Google Scholar ] [ CrossRef ] [ PubMed ]
- Sage, R.F. A portrait of the C4 photosynthetic family on the 50th anniversary of its discovery: Species number, evolutionary lineages, and Hall of Fame. J. Exp. Bot. 2017 , 68 , e11–e28. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
- Bianconi, M.E.; Dunning, L.T.; Moreno-Villena, J.J.; Osborne, C.P.; Christin, P.A. Gene duplication and dosage effects during the early emergence of C 4 photosynthesis in the grass genus Alloteropsis . J. Exp. Bot. 2018 , 69 , 1967–1980. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
- Williams, B.P.; Aubry, S.; Hibberd, J.M. Molecular evolution of genes recruited into C 4 photosynthesis. Trends Plant Sci. 2012 , 17 , 213–220. [ Google Scholar ] [ CrossRef ]
- von Caemmerer, S.; Furbank, R.T. Strategies for improving C 4 photosynthesis. Curr. Opin. Plant Biol. 2016 , 31 , 125–134. [ Google Scholar ] [ CrossRef ] [ Green Version ]
- Nishiyama, Y.; Murata, N. Revised scheme for the mechanism of photoinhibition and its application to enhance the abiotic stress tolerance of the photosynthetic machinery. Appl. Microbiol. Biotechnol. 2014 , 98 , 8777–8796. [ Google Scholar ] [ CrossRef ]
- Ghannoum, O. C 4 photosynthesis and water stress. Ann. Bot. 2009 , 103 , 635–644. [ Google Scholar ] [ CrossRef ] [ Green Version ]
- Osborne, C.P.; Freckleton, R.P. Ecological selection pressures for C 4 photosynthesis in the grasses. Proc. R. Soc. B Biol. Sci. 2009 , 276 , 1753–1760. [ Google Scholar ] [ CrossRef ] [ Green Version ]
- Guidi, L.; Lo Piccolo, E.; Landi, M. Chlorophyll fluorescence, photoinhibition and abiotic stress: Does it make any difference the fact to be a C 3 or C 4 species? Front. Plant Sci. 2019 , 10 , 174. [ Google Scholar ] [ CrossRef ]
- Ishikawa, N.; Takabayashi, A.; Noguchi, K.; Tazoe, Y.; Yamamoto, H.; Von Caemmerer, S.; Sato, F.; Endo, T. NDH-Mediated Cyclic Electron Flow Around Photosystem I is Crucial for C 4 Photosynthesis. Plant Cell Physiol. 2016 , 57 , 2020–2028. [ Google Scholar ] [ CrossRef ] [ Green Version ]
- Nakamura, N.; Iwano, M.; Havaux, M.; Yokota, A.; Munekage, Y.N. Promotion of cyclic electron transport around photosystem I during the evolution of NADP-malic enzyme-type C 4 photosynthesis in the genus Flaveria . New Phytol. 2013 , 199 , 832–842. [ Google Scholar ] [ CrossRef ] [ PubMed ]
- Hertle, A.P.; Blunder, T.; Wunder, T.; Pesaresi, P.; Pribil, M.; Armbruster, U.; Leister, D. PGRL1 is the elusive ferredoxin-plastoquinone reductase in photosynthetic cyclic electron flow. Mol. Cell 2013 , 49 , 511–523. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
- Darie, C.C.; Biniossek, M.L.; Winter, V.; Mutschler, B.; Haehnel, W. Isolation and structural characterization of the Ndh complex from mesophyll and bundle sheath chloroplasts of Zea mays . FEBS J. 2005 , 272 , 2705–2716. [ Google Scholar ] [ CrossRef ] [ PubMed ]
- Kubicki, A.; Funk, E.; Westhoff, P.; Steinmüller, K. Differential expression of plastome-encoded ndh genes in mesophyll and bundle-sheath chloroplasts of the C 4 plant Sorghum bicolor indicates that the complex I-homologous NAD(P)H-plastoquinone oxidoreductase is involved in cyclic electron transport. Planta 1996 , 199 , 276–281. [ Google Scholar ] [ CrossRef ]
- Majeran, W.; Zybailov, B.; Ytterberg, A.J.; Dunsmore, J.; Sun, Q.; van Wijk, K.J. Consequences of C 4 differentiation for chloroplast membrane proteomes in maize mesophyll and bundle sheath cells. Mol. Cell. Proteom. 2008 , 7 , 1609–1638. [ Google Scholar ] [ CrossRef ] [ Green Version ]
- Munekage, Y.N.; Taniguchi, Y.Y. Promotion of Cyclic Electron Transport around Photosystem I with the Development of C 4 Photosynthesis. Plant Cell Physiol. 2016 , 57 , 897–903. [ Google Scholar ] [ CrossRef ] [ Green Version ]
- Takabayashi, A.; Kishine, M.; Asada, K.; Endo, T.; Sato, F. Differential use of two cyclic electron flows around photosystem I for driving CO 2 -concentration mechanism in C 4 photosynthesis. Proc. Natl. Acad. Sci. USA 2005 , 102 , 16898–16903. [ Google Scholar ] [ CrossRef ] [ Green Version ]
- Sazanov, L.A.; Burrows, P.A.; Nixon, P.J. The chloroplast Ndh complex mediates the dark reduction of the plastoquinone pool in response to heat stress in tobacco leaves. FEBS Lett. 1998 , 429 , 115–118. [ Google Scholar ] [ CrossRef ] [ Green Version ]
- Rogowski, P.; Wasilewska-Dębowska, W.; Urban, A.; Romanowska, E. Maize bundle sheath chloroplasts—A unique model of permanent State 2. Environ. Exp. Bot. 2018 , 155 , 321–331. [ Google Scholar ] [ CrossRef ]
- Romanowska, E.; Powikrowska, M.; Zienkiewicz, M.; Drozak, A.; Pokorska, B. High light induced accumulation of two isoforms of the CF1 α-subunit in mesophyll and bundle sheath chloroplasts of C 4 plants. Acta Biochim. Pol. 2008 , 55 , 175–182. [ Google Scholar ] [ CrossRef ]
- Shikanai, T. Chloroplast NDH: A different enzyme with a structure similar to that of respiratory NADH dehydrogenase. Biochim. Biophys. Acta Bioenerg. 2016 , 1857 , 1015–1022. [ Google Scholar ] [ CrossRef ] [ PubMed ]
- Krieger-Liszkay, A.; Feilke, K. The Dual Role of the Plastid Terminal Oxidase PTOX: Between a Protective and a Pro-oxidant Function. Front. Plant Sci. 2016 , 6 , 1147:1–1147:3. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
- Lennon, A.M.; Prommeenate, P.; Nixon, P.J. Location, expression and orientation of the putative chlororespiratory enzymes, Ndh and IMMUTANS, in higher-plant plastids. Planta 2003 , 218 , 254–260. [ Google Scholar ] [ CrossRef ] [ PubMed ]
- Streb, P.; Josse, E.M.; Gallouët, E.; Baptist, F.; Kuntz, M.; Cornic, G. Evidence for alternative electron sinks to photosynthetic carbon assimilation in the high mountain plant species Ranunculus glacialis . Plant. Cell Environ. 2005 , 28 , 1123–1135. [ Google Scholar ] [ CrossRef ]
- Majeran, W.; van Wijk, K.J. Cell-type-specific differentiation of chloroplasts in C4 plants. Trends Plant Sci. 2009 , 14 , 100–109. [ Google Scholar ] [ CrossRef ]
- Essemine, J.; Lyu, M.J.A.; Qu, M.; Perveen, S.; Khan, N.; Song, Q.; Chen, G.; Zhu, X.G. Contrasting Responses of Plastid Terminal Oxidase Activity Under Salt Stress in Two C4 Species With Different Salt Tolerance. Front. Plant Sci. 2020 , 11 , 1009. [ Google Scholar ] [ CrossRef ]
- Rumeau, D.; Peltier, G.; Cournac, L. Chlororespiration and cyclic electron flow around PSI during photosynthesis and plant stress response. Plant. Cell Environ. 2007 , 30 , 1041–1051. [ Google Scholar ] [ CrossRef ]
- Leong, T.Y.; Anderson, J.M. Adaptation of the thylakoid membranes of pea chloroplasts to light intensities. I. Study on the distribution of chlorophyll-protein complexes. Photosynth. Res. 1984 , 5 , 105–115. [ Google Scholar ] [ CrossRef ]
- Drozak, A.; Romanowska, E. Acclimation of mesophyll and bundle sheath chloroplasts of maize to different irradiances during growth. Biochim. Biophys. Acta Bioenerg. 2006 , 1757 , 1539–1546. [ Google Scholar ] [ CrossRef ] [ Green Version ]
- Pfannschmidt, T.; Nilsson, A.; Allen, J.F. Photosynthetic control of chloroplast gene expression. Nature 1999 , 397 , 625–628. [ Google Scholar ] [ CrossRef ]
- Urban, A.; Rogowski, P.; Wasilewska-Dębowska, W.; Romanowska, E. Effect of light on the rearrangements of PSI super-and megacomplexes in the non-appressed thylakoid domains of maize mesophyll chloroplasts. Plant Sci. 2020 , 301 , 110655. [ Google Scholar ] [ CrossRef ] [ PubMed ]
- Murata, N.; Takahashi, S.; Nishiyama, Y.; Allakhverdiev, S.I. Photoinhibition of photosystem II under environmental stress. Biochim. Biophys. Acta Bioenerg. 2007 , 1767 , 414–421. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
- Demmig-Adams, B.; Cohu, C.M.; Muller, O.; Adams, W.W. Modulation of photosynthetic energy conversion efficiency in nature: From seconds to seasons. Photosynth. Res. 2012 , 113 , 75–88. [ Google Scholar ] [ CrossRef ] [ PubMed ]
- Kangasjärvi, S.; Neukermans, J.; Li, S.; Aro, E.M.; Noctor, G. Photosynthesis, photorespiration, and light signalling in defence responses. J. Exp. Bot. 2012 , 63 , 1619–1636. [ Google Scholar ] [ CrossRef ] [ Green Version ]
- Tikkanen, M.; Grebe, S. Switching off photoprotection of photosystem I—A novel tool for gradual PSI photoinhibition. Physiol. Plant. 2018 , 162 , 156–161. [ Google Scholar ] [ CrossRef ] [ PubMed ]
- Tikkanen, M.; Mekala, N.R.; Aro, E.M. Photosystem II photoinhibition-repair cycle protects Photosystem I from irreversible damage. Biochim. Biophys. Acta 2014 , 1837 , 210–215. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
- Ballottari, M.; Alcocer, M.J.P.; D’Andrea, C.; Viola, D.; Ahn, T.K.; Petrozza, A.; Polli, D.; Fleming, G.R.; Cerullo, G.; Bassi, R. Regulation of photosystem I light harvesting by zeaxanthin. Proc. Natl. Acad. Sci. USA 2014 , 111 , E2431–E2438. [ Google Scholar ] [ CrossRef ] [ Green Version ]
- Lima-Melo, Y.; Alencar, V.T.C.B.; Lobo, A.K.M.; Sousa, R.H.V.; Tikkanen, M.; Aro, E.M.; Silveira, J.A.G.; Gollan, P.J. Photoinhibition of Photosystem I Provides Oxidative Protection During Imbalanced Photosynthetic Electron Transport in Arabidopsis thaliana . Front. Plant Sci. 2019 , 10 , 916:1–916:13. [ Google Scholar ] [ CrossRef ] [ Green Version ]
- Pokorska, B.; Zienkiewicz, M.; Powikrowska, M.; Drozak, A.; Romanowska, E. Differential turnover of the photosystem II reaction centre D1 protein in mesophyll and bundle sheath chloroplasts of maize. Biochim. Biophys. Acta Bioenerg. 2009 , 1787 , 1161–1169. [ Google Scholar ] [ CrossRef ] [ Green Version ]
- Rintamäki, E.; Kettunen, R.; Aro, E.M. Differential D1 Dephosphorylation in Functional and Photodamaged Photosystem II Centers: Dephosphorylation is a prerequisite for degradation of damaged D1. J. Biol. Chem. 1996 , 271 , 14870–14875. [ Google Scholar ] [ CrossRef ] [ Green Version ]
- Zhang, L.; Paakkarinen, V.; Van Wijk, K.J.; Aro, E.M. Biogenesis of the chloroplast-encoded D1 protein: Regulation of translation elongation, insertion, and assembly into photosystem II. Plant Cell 2000 , 12 , 1769–1781. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
- Tikkanen, M.; Grieco, M.; Kangasjärvi, S.; Aro, E.M. Thylakoid Protein Phosphorylation in Higher Plant Chloroplasts Optimizes Electron Transfer under Fluctuating Light. Plant Physiol. 2010 , 152 , 723–735. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
- Jahns, P.; Holzwarth, A.R. The role of the xanthophyll cycle and of lutein in photoprotection of photosystem II. Biochim. Biophys. Acta 2012 , 1817 , 182–193. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
- Zienkiewicz, M.; Krupnik, T.; Drozak, A.; Golke, A.; Romanowska, E. Chloramphenicol acetyltransferase- a new selectable marker in stable nuclear transformation of the red alga Cyanidioschyzon merolae . Protoplasma 2017 , 254 , 587–596. [ Google Scholar ] [ CrossRef ]
- Sharma, P.K.; Hall, D.O. Changes in Carotenoid Composition and Photosynthesis in Sorghum under High Light and Salt Stresses. J. Plant Physiol. 1992 , 140 , 661–666. [ Google Scholar ] [ CrossRef ]
- Bergantino, E.; Segalla, A.; Brunetta, A.; Teardo, E.; Rigoni, F.; Giacometti, G.M.; Szabò, I. Light- and pH-dependent structural changes in the PsbS subunit of photosystem II. Proc. Natl. Acad. Sci. USA 2003 , 100 , 15265–15270. [ Google Scholar ] [ CrossRef ] [ Green Version ]
Process Taking Place in Chloroplasts | C3 Plants | C4 Plants |
---|---|---|
Xanthophyll cycle and heat dissipation | Typical, occurring with the zeaxanthin and PsbS protein [ ]. | |
State transitions and LHCII phosphorylation | Function of state 1 and state 2, depending on phosphorylation of the LHCII antenna [ ]. | Permanent state 2 in agranal BS maize (NADP-ME) chloroplasts. LHCII in phosphorylated form, regardless of the condition [ ]. |
Photoinhibition and phosphorylation of D1 protein | Damaged D1 is directed to the thylakoid stroma, dephosphorylated, and then degraded. | D1 degradation is faster in the BS chloroplast of maize [ ]. Photodamage of some PSII pools for protection against PSI excess [ ]. |
Cyclic electron transport components | Lower ATP demand resulting from metabolism. | Elevation of the CET ad alternative CET pathway with NDH complex for higher efficiency of ATP production [ , , ]. |
PTOX functioning and chlororespiration | Minor importance, activity mainly under stressful conditions. | High amount and activity in maize BS chloroplasts for better protection against ROS formation during elevated cyclic electron transport [ ]. |
Changes in antenna and reaction centers amount | Higher content of LHCII antenna in low light intensities. Higher content of reaction centers at high light intensities [ ]. | |
Additional mechanism(s) | No data available. | Formation of megacomplexes in maize mesophyll chloroplasts [ ]. |
MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations. |
Share and Cite
Wasilewska-Dębowska, W.; Zienkiewicz, M.; Drozak, A. How Light Reactions of Photosynthesis in C4 Plants Are Optimized and Protected under High Light Conditions. Int. J. Mol. Sci. 2022 , 23 , 3626. https://doi.org/10.3390/ijms23073626
Wasilewska-Dębowska W, Zienkiewicz M, Drozak A. How Light Reactions of Photosynthesis in C4 Plants Are Optimized and Protected under High Light Conditions. International Journal of Molecular Sciences . 2022; 23(7):3626. https://doi.org/10.3390/ijms23073626
Wasilewska-Dębowska, Wioleta, Maksymilian Zienkiewicz, and Anna Drozak. 2022. "How Light Reactions of Photosynthesis in C4 Plants Are Optimized and Protected under High Light Conditions" International Journal of Molecular Sciences 23, no. 7: 3626. https://doi.org/10.3390/ijms23073626
Article Metrics
Article access statistics, further information, mdpi initiatives, follow mdpi.
Subscribe to receive issue release notifications and newsletters from MDPI journals
An official website of the United States government
The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.
The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.
- Publications
- Account settings
- My Bibliography
- Collections
- Citation manager
Save citation to file
Email citation, add to collections.
- Create a new collection
- Add to an existing collection
Add to My Bibliography
Your saved search, create a file for external citation management software, your rss feed.
- Search in PubMed
- Search in NLM Catalog
- Add to Search
The effect of light quality on plant physiology, photosynthetic, and stress response in Arabidopsis thaliana leaves
Affiliations.
- 1 Department of Bioresource Engineering, McGill University-Macdonald Campus, Sainte-Anne-de-Bellevue, Quebec, Canada.
- 2 Department of Plant Science, McGill University-Macdonald Campus, Sainte-Anne-de-Bellevue, Quebec, Canada.
- PMID: 33661984
- PMCID: PMC7932170
- DOI: 10.1371/journal.pone.0247380
The impacts of wavelengths in 500-600 nm on plant response and their underlying mechanisms remain elusive and required further investigation. Here, we investigated the effect of light quality on leaf area growth, biomass, pigments content, and net photosynthetic rate (Pn) across three Arabidopsis thaliana accessions, along with changes in transcription, photosynthates content, and antioxidative enzyme activity. Eleven-leaves plants were treated with BL; 450 nm, AL; 595 nm, RL; 650 nm, and FL; 400-700 nm as control. RL significantly increased leaf area growth, biomass, and promoted Pn. BL increased leaf area growth, carotenoid and anthocyanin content. AL significantly reduced leaf area growth and biomass, while Pn remained unaffected. Petiole elongation was further observed across accessions under AL. To explore the underlying mechanisms under AL, expression of key marker genes involved in light-responsive photosynthetic reaction, enzymatic activity of antioxidants, and content of photosynthates were monitored in Col-0 under AL, RL (as contrast), and FL (as control). AL induced transcription of GSH2 and PSBA, while downregulated NPQ1 and FNR2. Photosynthates, including proteins and starches, showed lower content under AL. SOD and APX showed enhanced enzymatic activity under AL. These results provide insight into physiological and photosynthetic responses to light quality, in addition to identifying putative protective-mechanisms that may be induced to cope with lighting-stress in order to enhance plant stress tolerance.
PubMed Disclaimer
Conflict of interest statement
The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.
Fig 1. Effect of BL, AL, RL…
Fig 1. Effect of BL, AL, RL and FL on morphology, leaf area growth, biomass,…
Fig 2. Flow diagram of study design…
Fig 2. Flow diagram of study design to investigate the potential induction of stress response…
Fig 3. A schematic model of light-responsive…
Fig 3. A schematic model of light-responsive photosynthetic process and effect of AL and RL…
Fig 4. Effect of AL and RL…
Fig 4. Effect of AL and RL on photosynthates content in A . thaliana Col-0.
Fig 5. Effect of AL and RL…
Fig 5. Effect of AL and RL on SOD and APX activity in A .…
Similar articles
- Impact of high irradiance and UV-B on the photosynthetic activity, pro-/antioxidant balance and expression of light-activated genes in Arabidopsis thaliana hy4 mutants grown under blue light. Kreslavski VD, Khudyakova AY, Strokina VV, Shirshikova GN, Pashkovskiy PP, Balakhnina TI, Kosobryukhov AA, Kuznetsov VV, Allakhverdiev SI. Kreslavski VD, et al. Plant Physiol Biochem. 2021 Oct;167:153-162. doi: 10.1016/j.plaphy.2021.07.030. Epub 2021 Jul 26. Plant Physiol Biochem. 2021. PMID: 34358729
- Enhancement of leaf photosynthetic capacity through increased stomatal density in Arabidopsis. Tanaka Y, Sugano SS, Shimada T, Hara-Nishimura I. Tanaka Y, et al. New Phytol. 2013 May;198(3):757-764. doi: 10.1111/nph.12186. Epub 2013 Feb 25. New Phytol. 2013. PMID: 23432385
- Chloroplast Accumulation Response Enhances Leaf Photosynthesis and Plant Biomass Production. Gotoh E, Suetsugu N, Yamori W, Ishishita K, Kiyabu R, Fukuda M, Higa T, Shirouchi B, Wada M. Gotoh E, et al. Plant Physiol. 2018 Nov;178(3):1358-1369. doi: 10.1104/pp.18.00484. Epub 2018 Sep 28. Plant Physiol. 2018. PMID: 30266749 Free PMC article.
- Photomorphogenesis of leaves: shade-avoidance and differentiation of sun and shade leaves. Kim GT, Yano S, Kozuka T, Tsukaya H. Kim GT, et al. Photochem Photobiol Sci. 2005 Sep;4(9):770-4. doi: 10.1039/b418440h. Epub 2005 May 4. Photochem Photobiol Sci. 2005. PMID: 16121290 Review.
- [Photoprotective mechanisms of leaf anthocyanins: research progress]. Wang LZ, Hu YB, Zhang HH, Xu N, Zhang XL, Sun GY. Wang LZ, et al. Ying Yong Sheng Tai Xue Bao. 2012 Mar;23(3):835-41. Ying Yong Sheng Tai Xue Bao. 2012. PMID: 22720633 Review. Chinese.
- Abiotic Stresses in Plants: From Molecules to Environment. Bartas M. Bartas M. Int J Mol Sci. 2024 Jul 24;25(15):8072. doi: 10.3390/ijms25158072. Int J Mol Sci. 2024. PMID: 39125642 Free PMC article.
- Alfalfa MsbHLH115 confers tolerance to cadmium stress through activating the iron deficiency response in Arabidopsis thaliana . Zhang M, Gao JY, Dong SC, Chang MH, Zhu JX, Guo DL, Guo CH, Bi YD. Zhang M, et al. Front Plant Sci. 2024 Feb 12;15:1358673. doi: 10.3389/fpls.2024.1358673. eCollection 2024. Front Plant Sci. 2024. PMID: 38410731 Free PMC article.
- Shade Effect on Phenology, Fruit Yield, and Phenolic Content of Two Wild Blueberry Species in Northwestern Ontario, Canada. Dyukaryeva V, Mallik AU. Dyukaryeva V, et al. Plants (Basel). 2023 Dec 7;12(24):4099. doi: 10.3390/plants12244099. Plants (Basel). 2023. PMID: 38140426 Free PMC article.
- White LED Lighting Increases the Root Productivity of Panax ginseng C. A. Meyer in a Hydroponic Cultivation System of a Plant Factory. Kim SH, Park JH, Kim EJ, Lee JM, Park JW, Kim YS, Kim GR, Lee JS, Lee EP, You YH. Kim SH, et al. Biology (Basel). 2023 Jul 26;12(8):1052. doi: 10.3390/biology12081052. Biology (Basel). 2023. PMID: 37626938 Free PMC article.
- The Role of Light Quality in Regulating Early Seedling Development. Wei Y, Wang S, Yu D. Wei Y, et al. Plants (Basel). 2023 Jul 24;12(14):2746. doi: 10.3390/plants12142746. Plants (Basel). 2023. PMID: 37514360 Free PMC article. Review.
- Smith H., "Phytochromes and light signal perception by plants—an emerging synthesis," Nature, vol. 407, no. 6804, pp. 585–91, October 5 2000, 10.1038/35036500 - DOI - PubMed
- Ceccarelli E. A., Arakaki A. K., Cortez N., and Carrillo N., "Functional plasticity and catalytic efficiency in plant and bacterial ferredoxin-NADP (H) reductases," Biochimica et Biophysica Acta (BBA)-Proteins and Proteomics, vol. 1698, no. 2, pp. 155–165, 2004. - PubMed
- McCree K., "The action spectrum, absorptance and quantum yield of photosynthesis in crop plants," Agricultural Meteorology, vol. 9, pp. 191–216, 1972.
- Belkov V., Garnik E. Y., and Konstantinov Y. M., "Mechanism of plant adaptation to changing illumination by rearrangements of their photosynthetic apparatus," Current Challenges in Plant Genetics, Genomics, Bioinformatics, and Biotechnology, vol. 24, p. 101, 2019.
- Inada K., "Action spectra for photosynthesis in higher plants," Plant and Cell Physiology, vol. 17, no. 2, pp. 355–365, 1976.
Publication types
- Search in MeSH
Grants and funding
Linkout - more resources, full text sources.
- Europe PubMed Central
- PubMed Central
- Public Library of Science
Other Literature Sources
- scite Smart Citations
Molecular Biology Databases
- The Arabidopsis Information Resource
Miscellaneous
- NCI CPTAC Assay Portal
- Citation Manager
NCBI Literature Resources
MeSH PMC Bookshelf Disclaimer
The PubMed wordmark and PubMed logo are registered trademarks of the U.S. Department of Health and Human Services (HHS). Unauthorized use of these marks is strictly prohibited.
- Search Menu
- Sign in through your institution
- Volume 16, Issue 5, October 2024 (In Progress)
- Volume 16, Issue 4, July 2024
- Advance articles
- Editor's Choice
- Special Issues
- Author Guidelines
- Article Metrics
- Review Procedures and Criteria
- Open Access Policy
- Publication Charges
- Submission Site
- Call for Papers
- Why Publish?
- About AoB PLANTS
- Impact Factor
- Editorial Board
- About the Annals of Botany Company
- Advertising and Corporate Services
- Journals Career Network
- Self-Archiving Policy
- Journals on Oxford Academic
- Books on Oxford Academic
Article Contents
Introduction, materials and methods, conclusions, sources of funding, contributions by the authors, conflict of interest, supporting information, acknowledgements, literature cited.
- < Previous
Effects of growth under different light spectra on the subsequent high light tolerance in rose plants
- Article contents
- Figures & tables
- Supplementary Data
Leyla Bayat, Mostafa Arab, Sasan Aliniaeifard, Mehdi Seif, Oksana Lastochkina, Tao Li, Effects of growth under different light spectra on the subsequent high light tolerance in rose plants, AoB PLANTS , Volume 10, Issue 5, October 2018, ply052, https://doi.org/10.1093/aobpla/ply052
- Permissions Icon Permissions
Photosynthesis is defined as a light-dependent process; however, it is negatively influenced by high light (HL) intensities. To investigate whether the memory of growth under monochromatic or combinational lights can influence plant responses to HL, rose plants were grown under different light spectra [including red (R), blue (B), 70:30 % red:blue (RB) and white (W)] and were exposed to HL (1500 μmol m −2 s −1 ) for 12 h. Polyphasic chlorophyll a fluorescence (OJIP) transients revealed that although monochromatic R- and B-grown plants performed well under control conditions, the functionality of their electron transport system was more sensitive to HL than that of the RB- and W-grown plants. Before exposure to HL, the highest anthocyanin concentration was observed in R- and B-grown plants, while exposure to HL reduced anthocyanin concentration in both R- and B-grown plants. Ascorbate peroxidase and catalase activities decreased, while superoxide dismutase activity was increased after exposure to HL. This caused an increase in H 2 O 2 concentration and malondialdehyde content following HL exposure. Soluble carbohydrates were decreased by exposure to HL, and this decrease was more emphasized in R- and B-grown plants. In conclusion, growing plants under monochromatic light reduced the plants ability to cope with HL stress.
Light is the original source of energy for plant photosynthesis and growth. A wide range of signals and information for morphogenesis and many other physiological processes is triggered by light ( Chen et al. 2004 ). Different characteristics of light such as spectral composition (wavelengths), intensity, duration and direction can influence plant growth and development. The photosynthesis process is also sensitive to all aspects of lighting environments.
The rapid development of lighting technologies using light-emitting diodes (LEDs) has caused an increase in the application of this technology for lighting in closed horticultural systems ( Kozai et al. 2015 ). Light-emitting diodes are also attractive because of their high radiant efficiency, long lifetimes, small size, low temperature, narrow spectrum and physical robustness ( Tennessen et al. 1994 ; Kim et al. 2004 ; Breive et al. 2005 ; Morrow 2008 ). The application of LEDs in horticulture makes it possible to use certain wavelengths to study particular plant responses. For instance, intercrop lighting using LEDs is nowadays used to promote photosynthesis of the middle and lower leaves. However, research on the effect of spectral wavelengths on plant growth and development is still in progress.
Plant responses to light differ based on the lighting environment, season, genotype, cultivation practices and many others ( Kozai 2016 ). Although light is the energy source for photosynthesis, it can simultaneously function as a stress factor. Under high light (HL) intensity conditions or when plants are exposed to other abiotic stresses (e.g. drought), the energy supply (ATP) and reducing power (NADPH) through photosystems (and by involvement of electron transport chain) exceed the demand for metabolic processes in carbon-fixing reactions ( Miyake et al. 2009 ; Gu et al. 2017 ). The accumulation of reactive oxygen species (ROS) is the result of a disturbance between supply and demand for electron transport end products. Depending on the rate of accumulation, ROS can have dual effects on plant responses. At low concentrations, ROS act as a signal to induce defence responses, while at high concentrations ROS are toxic and induce lipid peroxidation in the cell membranes and cause oxidative damage to cellular components ( Dat et al. 2000 ; Vranová et al. 2002 ). Consequently, the functionality of photosynthesis decreases under stressful conditions. Photosynthesis suppression due to exposure to HL intensities is known as photoinhibition. Damage caused by the photoinhibition process can take place in all components of the photosynthetic machinery. Among them, the photosystem II (PSII) complex of the electron transport chain is considered the primary target of photoinhibition ( Vass 2012 ; Roach and Krieger-Liszkay 2014 ). However, photosynthesis as a fine-tuned process employs multiple mechanisms to cope with photoinhibition. Dissipation of excess energy in the form of heat is the main protective strategy to get rid of the damage induced by HL intensities ( Müller et al. 2001 ). Thermal dissipation is known as non-photochemical quenching (NPQ) of Chl fluorescence. Non-photochemical quenching encompasses several strategies including: energy-dependent quenching by involvement of the xanthophyll cycle, conformational changes in light harvesting complex II (LHCII), which are known as state transitions, and photoinhibition resulting in reduction of quantum yield as a consequence of light-induced damage ( Müller et al. 2001 ; Zhao et al. 2017 ).
Light is absorbed by plant pigments. Chl a and b are the main photosynthetic pigments in plants. They mainly absorb blue and red wavelengths of the light spectrum. Carotenoids with an absorption spectrum between 350 and 500 nm are also found in all chlorophyll-based photosynthesis systems. They contribute to light absorption in the antenna system. Furthermore, carotenoids help the plant to dissipate excess energy as heat to protect the photosynthesis apparatus from HL intensities ( Sharma and Hall 1993 ). Anthocyanins are also involved in the protection of photosynthetic apparatus from damage due to HL intensity ( Petrella et al. 2016 ).
Roses are the most famous ornamental plants worldwide. The growth and development as well as post-harvest quality of rose plants are affected by pre-harvest environmental factors such as relative humidity (RH) and light ( Fanourakis et al. 2013 ). Issues related to light have attracted much attention for rose cultivation during the last few decades. In some places (especially at high latitudes), supplementary lighting is used to compensate sun light limitations. In other places exposure to HL intensities during the summer negatively influences rose production. In those places shading the rose plants is a common practice during summer time. Furthermore, in many places plants are only exposed to HL conditions for a short duration. However, both shading and HL can negatively influence photosynthesis ( García Victoria et al. 2011 ).
Each spectral band of light can induce certain responses in plants. For instance, red and blue lights mainly contain the range of wavelengths necessary for electron excitation in the photosynthetic apparatus ( Taiz and Zeiger 2002 ). Wavelengths in the range of blue and UV cause the accumulation of carotenoids and anthocyanins in the leaves ( Li and Kubota 2009 ; Carvalho et al. 2016 ), and red light induces accumulation of carbohydrates in the leaf ( Sæbø et al. 1995 ). Since each band of light wavelengths can induce certain mechanisms and responses in plants, which can affect subsequent plant responses to environmental stresses, in the current study, we hypothesize that protective mechanisms against HL stress would be up-regulated/down-regulated through growing plants under certain light spectra which can influence plant responses to HL intensity afterwards. Furthermore, since limited wavelengths can induce certain protective mechanisms against HL, plants grown under combined wavelengths have more capabilities to tolerate HL intensities.
Cryptochromes and phototropins are the B light acceptors, whereas phytochromes are more sensitive acceptors for R light ( Whitelam and Halliday 2008 ). These photoreceptors acquire information from the light environment and use that information to modulate cellular processes ( Smith 1982 ). Therefore, different signal transduction pathways are involved in the up- or down-regulation of certain metabolic pathways by the information gained from specific photoreceptor of certain wavelength. The aims of the current study were to investigate (i) the structure and function of photosynthetic apparatus in plants (grown under different light spectra) exposed to HL intensity, and (ii) the mechanisms involved in HL tolerance in rose plants grown under different light spectra.
It has been shown that the energy flow and information related to the structure and function of the photosynthetic apparatus can be analysed through Chl fluorescence data. Although chlorophyll fluorescence parameters obtained by the saturation pulse method in light-adapted leaves give valuable information regarding the energy flow across thylakoid membranes, these types of measurement are time-consuming and are not fully ideal for assessing fluorescence parameters in practical conditions ( Brestic and Zivcak 2013 ). The non-destructive analysis of polyphasic fast chlorophyll transient by the so-called OJIP test was developed for quick evaluation of biophysical aspects of photosynthesis especially under stress conditions ( Strasser 1995 ; Mathur et al. 2013 ). This test relies on energy flow in thylakoid membranes, which provides detailed information about the biophysics of the photosynthetic system through measurement of fluorescence signals ( Kalaji et al. 2017 ). The OJIP test has been successfully used for studying the photosynthetic apparatus of different plant species under abiotic stress conditions ( Panda et al. 2008 ; Martinazzo et al. 2012 ; Kalaji et al. 2014 ; Rapacz et al. 2015 ; Kalaji et al. 2016 ). Therefore, this test was used in the current study to detect the negative effects of HL stress on the photosynthetic apparatus of the rose plants.
Plant material and growth conditions
Rose plants ( Rosa hybrida cv. ‘Avalanche’) were propagated from cuttings in March 2017 in a perlite medium. Rooting was achieved following 4–5 weeks of planting. The rooted cuttings (10–15 cm long with two emerging leaves) were transplanted into 15 cm diameter plastic pots containing a mixture of cocopite and perlite (70/30 % by volume). The plants were watered daily for 3 days after transplantation, and thereafter, they were irrigated with half-strength Hoagland solution. Homogeneous plants were divided into four growth chambers (four plant per each growth chamber) with the same climatic conditions (temperature: 27 ± 3 °C; RH: 50 ± 5 %; photoperiod: 12-h light/12-h dark between 0800 and 2000 h). Plants were subjected to four different light spectra: white (W), blue (B), red (R) and 70 % red + 30 % blue (RB) provided by LED production modules (24-W, Iran Grow Light Co., Iran) at a photosynthetic photon flux density (PPFD) of 250 ± 10 μmol m −2 s −1 . The W spectrum consisted of 41 % in the range of B and 18 % in the range of R. Photosynthetic photon flux density intensities and light spectra were monitored using a light meter (Sekonic C-7000, Japan). The relative spectra of the light treatments are shown in Fig. 1 . Four weeks after growth at different light spectra, all plants were exposed to a HL intensity (1500 μmol m −2 s −1 ) for 12 h (between 0800 and 2000 h) ( Fig. 2 ). To provide the same temperature during plant exposure to HL intensity as the growth chamber temperature, plants were illuminated with LED production modules (Iran Grow Light Co., Iran) in the range of 400–700 nm in a temperature-controlled room at 21 °C and 50 % RH. Under this condition, the temperature was fixed to 27 °C under HL intensity (same as the temperature in the growth chambers). Since such a HL intensity caused an increase in the temperature of growth chamber (27 °C), to maintain the same temperature before and after HL intensity same temperature was also kept inside the growth chambers during the growth of plants. The module consisted of 320 LEDs equipped with three fan ventilators to reduce the heat produced in the module. Measurement of the parameters was performed at two time points: first, at the end of plant growth under different light spectra with PPFD of 250 μmol m −2 s −1 , and second, at the end of plant exposure to HL intensity (1500 μmol m −2 s −1 ).
Light spectra of the blue (B), red (R), red and blue (RB) and white (W) lighting environments measured at plant level in the growth chambers.
Representative images showing plants (A) that were grown for 3 weeks under different light spectra [blue (B), red (R), white (W) and red and blue (RB)], growth chambers that were used for growing plants under 250 (B) and 1500 (C) µmol m −2 s −1 PPFD.
Chl fluorescence and OJIP test measurements
The youngest fully developed leaves were used for measuring the maximum quantum efficiency of PSII ( F v / F m ) with the Handy Fluorcam FC 1000-H (Photon Systems Instruments, PSI, Czech Republic). Intact leaves attached to the plants were dark-adapted for 20 min. After dark adaptation, intact plants were immediately used to measure F v / F m . The Fluorcam consisting of a CCD camera and four fixed LED panels, one pair supplying the measuring pulses and the second pair providing actinic illumination and saturating flash, were used. F v / F m was calculated using a custom-made protocol ( Genty et al. 1989 ; Aliniaeifard et al. 2014 ; Aliniaeifard and van Meeteren 2014 ). Images were recorded during short measuring flashes in darkness. At the end of the short flashes, the samples were exposed to a saturating light pulse (3900 µmol m −2 s −1 ) that resulted in a transitory saturation of photochemistry and reduction of the primary quinone acceptor of PSII ( Genty et al. 1989 ). After reaching steady-state fluorescence, two successive series of fluorescence data were digitized and averaged, one during the short measuring flashes in darkness ( F 0 ), and the other during the saturating light flash ( F m ). From these two images, F v was calculated by the expression F v = F m − F 0 . The F v / F m was calculated using the ratio ( F m − F 0 )/ F m . Maximum fluorescence ( F m ′) was determined in light-adapted steady state and was used for calculation of NPQ according to the following equation:
The average values and standard deviation per image were calculated using Fluorcam software version 7.0.
The polyphasic Chl a fluorescence (OJIP) transients were measured using a Fluorpen FP 100-MAX (Photon Systems Instruments, Drasov, Czech Republic) on young fully expanded rose leaves after 20 min dark adaptation. OJIP was used to study different biophysical and phenomenological parameters related to PSII status ( Strasser 1995 ). The transient fluorescence measurement was induced by a saturating light of ~3000 μmol m −2 s −1 . All the measurements were performed on 20 min dark-adapted plants. The OJIP transients were done according to the JIP test ( Strasser et al. 2000b ). The following data from the original measurements were used after extraction by Fluorpen software: fluorescence intensities at 50 μs (F50μs, considered as the minimum fluorescence F 0 ), 2 ms (J-step, F J ), 60 ms (I-step, F I ) and maximum fluorescence ( F m ).
The performance index was calculated on the absorption basis (PI ABS ) and densities of QA − reducing PSII reaction centres at time 0 and time to reach maximum fluorescence. The yield ratios, including the probability that a trapped exciton moves an electron in the electron transport chain beyond QA − (ψ o ), the quantum yield of electron transport (φ Eo ), the quantum yield of energy dissipation (φ Do ) and the maximum quantum yield of primary photochemistry (φ Po ), were also calculated based on the following equations:
From these data the following parameters were calculated: the specific energy fluxes per reaction centre (RC) for energy absorption (ABS/RC = M 0 ·(1/V J )·(1/φ Po )), (M 0 = TR 0 /RC-ET 0 /RC), trapped energy flux (TR 0 /RC = M 0 · (1/V J )), electron transport flux (ET 0 /RC = M 0 ·(1/V J )·ψ o ) and dissipated energy flux (DI 0 /RC = (ABS/RC) – (TR 0 /RC)).
Pigment measurements
The Chl and carotenoid contents of the leaves were measured according to the method described by Arnon (1949) . For measuring the anthocyanin content of the leaves, 1 g of leaf tissue was homogenized in 10 mL methanol and the extract was incubated at 4 °C in the dark overnight. The slurry was centrifuged (SIGMA-3K30) at 4000 g for 10 min. The anthocyanin in the supernatant was spectrophotometrically (Lambda 25 UV/VIS spectrometer) determined at 520 nm according to the method described by Wagner (1979) .
Determination of hydrogen peroxide content and lipid peroxidation
The level of lipid peroxidation in the leaf tissue was measured based on the malondialdehyde (MDA, a product of lipid peroxidation) content of the leaves. Malondialdehyde was determined by the thiobarbituric acid (TBA) reaction by minor modification of the method described by Heath and Packer (1968) . A 0.25 g leaf sample was homogenized in 5 mL of 0.1 % trichloroacetic acid (TCA). The homogenate was centrifuged at 10000 g for 5 min. Four millilitres of 20 % TCA containing 0.5 % TBA was added to a 1-mL aliquot of the supernatant. The mixture was heated at 95 °C for 30 min and then quickly cooled in an ice bath. After centrifuging at 10000 g for 10 min the absorbance of the supernatant at 532 nm was read and the value for the non-specific absorption at 600 nm was subtracted. The concentration of MDA was calculated using its extinction coefficient of 155 mM −1 cm −1 ( Heath and Packer 1968 ).
Hydrogen peroxide (H 2 O 2 ) content was spectrophotometrically measured after reaction with potassium iodide (KI). The reaction mixture contained 0.5 mL of 0.1 % TCA, leaf extract supernatant, 0.5 mL of 100 mm K-phosphate buffer and 2 mL reagent (1 M KI w/v in fresh double-distilled H 2 O). The blank contained 0.1 % TCA in the absence of leaf extract. The reaction was developed for 1 h in darkness and absorbance was measured at 390 nm. The amount of H 2 O 2 was calculated using a standard curve prepared with known concentrations of H 2 O 2 according to the method described by Patterson et al. (1984) .
Determination of carbohydrates
Young fully developed leaves (300 mg fresh weight [FW]) were collected from each replicate per treatment and were ground in liquid nitrogen, mixed with 7 mL of 70 % ethanol (w/v) for 5 min on ice and centrifuged at 6700 g for 10 min at 4 °C. After adding 200 mL of the supernatant to 1 mL of an anthrone solution (0.5 g anthrone, 250 mL 95 % H 2 SO 4 and 12.5 mL distilled water), the absorbance was spectrophotometrically recorded at 625 nm ( van Doorn 2012 ).
For starch quantification, 0.1 g of a fresh fully developed leaf was ground and sugars were extracted with 80 % ethanol and starch was solubilized using 52 % perchloric acid. Starch was colorimetrically determined at 630 nm using the sugar-anthrone-sulfuric acid reaction based on the method described by McCready et al. (1950) .
Determination of activities of antioxidant enzymes
Ascorbate peroxidase (APX) activity was determined by oxidation of ascorbic acid (AA) at 265 nm (ε = 13.7 mM −1 cm −1 ) by slight modification of the method described by Nakano and Asada (1981) . The reaction mixture contained 50 mM potassium phosphate buffer (pH 7.0), 5 mM AA, 0.5 mM H 2 O 2 and the enzyme extract. The reaction was started by adding H 2 O 2 . The rates were corrected for the non-enzymatic oxidation of AA by the inclusion of a reaction mixture without the enzyme extract (blind sample). The enzyme activity was expressed in μmol AA min −1 per g of fresh weight ( Nakano and Asada 1981 ).
For determination of catalase (CAT) activity, the decomposition of H 2 O 2 was recorded by the decrease in absorbance at 240 nm. Reaction mixture consisted of 1.5 mL 50 mM sodium phosphate buffer (pH 7.8), 0.3 mL 100 mM H 2 O 2 and 0.2 mL enzyme extract. One CAT unit was defined as the amount of enzyme necessary to decompose 1 mM min −1 H 2 O 2 . Therefore, the CAT activity was expressed as U g −1 FW min −1 ( Díaz-Vivancos et al. 2008 ).
Superoxide dismutase (SOD) activity was measured according to the method described by Dhindsa et al. (1981) . The method is based on the ability of SOD to inhibit the photochemical reduction of nitro blue tetrazolium (NBT). The reaction mixture contained 50 mM phosphate buffer (pH 7–8), 13 mM methionine, 75 µM NBT, 0.15 mM riboflavin, 0.1 mM EDTA and 0.50 mL enzyme extract. Riboflavin was the last item that was added. The reaction mixture was shaken and placed under two 15-W fluorescent lamps. The reaction was started by switching on the light and was allowed to run for 10 min. The reaction was stopped by turning off the light and the tubes were covered using a black cloth. The absorbance by the reaction mixture was spectrophotometrically recorded at 560 nm. A reaction mixture without exposure to the light was used as the control. No colour was observed in the control. Under assay conditions, one unit of SOD activity is expressed as the amount of enzyme that led to 50 % inhibition of NBT reduction ( Dhindsa et al. 1981 ).
Statistical analysis
Four plants were used as four replicates in each light treatment. Individual plants were taken as independent replicates. The measurements were done on two timescales: before and after exposure to HL. To make the environmental conditions of the two timescales similar, environmental conditions inside the chambers and between the chambers were kept as similar as possible before and after exposure to HL. To do this, the same temperature, RH and light intensity and duration were maintained during these two timescales in a climate-controlled room. Sampling was done on the same plants before and after exposure to HL. The data were subjected to two-way analysis of variance (ANOVA) and the Tukey test was used as a post-test. P > 0.05 was considered not significant. GraphPad Prism 7.01 for Windows (GraphPad Software, Inc., San Diego, CA) was used for the statistical analysis.
Polyphasic Chl a fluorescence (OJIP) transients
In the current study, the PSII activity was studied by calculation of different Chl fluorescence parameters in a leaf in the dark-adapted state. Investigating the fast chlorophyll fluorescence induction curve showed that all plants produced typical polyphasic curves with the basic OJIP steps ( Fig. 3 ). For all plants, the intensity of the fluorescence signal was increased from the initial fluorescence level ( F 0 ) to the intermediate steps (J and I) and then reached the maximum level ( F m ). Except for R light, exposure to HL caused a decline in fluorescence intensity in all steps of OJIP.
Fast chlorophyll fluorescence induction curve exhibited by leaves of rose plants exposed to different light spectra [blue (B), red (R), white (W) and red and blue (RB)] under 250 (C) and 1500 (HL) µmol m −2 s −1 PPFD.
Under control (C) conditions, the highest F 0 and F J were observed for W-grown plants and the lowest values were detected in R-grown plants ( Fig. 4A and B ). Under HL conditions, the highest F 0 , F J , F I and F P were observed for R-grown plants and their lowest values were detected in RB-grown plants ( Fig. 4A–D ). In all plants, exposure to HL resulted in a decrease in F v values; however, among the light treatments, the change in F v following exposure to HL was the lowest for W-grown plants ( Fig. 4E ). F v / F m also decreased following exposure to HL and this decrease was greatest for monochromatic R- and B-grown plants ( Fig. 4D ). The calculated parameters of the OJIP test were changed significantly due to exposure to different light spectra and HL, and the interactions were also significant for the calculated parameters [see Supporting Information—Table S1 ] . The calculated parameters for specific energy fluxes per reaction centre such as ABS/RC, TR 0 /RC and DI 0 /RC were increased following exposure to HL for all plants grown under different light spectra, except for the RB-grown plants ( Fig. 5A , B and D ). In the case of ABS/RC and DI 0 /RC, the highest difference between C and HL was observed in monochromatic R- and B-grown plants. There was no significant difference between C and HL for the RB-grown plants. ET 0 /RC was similar in the C and HL treatments for R- and B-grown plants, while it was increased in W- and RB-grown plants after exposure to HL ( Fig. 5C ).
Intensity of chlorophyll a fluorescence in the OJIP test including F 0 (A), F J (B), F I (C), F P (D), F v [E; ( F m − F 0 )] and F v / F m (F) from the fluorescence transient exhibited by leaves of rose plants grown under different light spectra [blue (B), red (R), white (W) and red and blue (RB)] under 250 (black bars) and 1500 (grey bars) µmol m −2 s −1 PPFD. Bars represent means ± SD.
Specific energy fluxes per reaction centre (RC) for energy absorption [A; (ABS/RC)], trapped energy flux [B; (TR 0 /RC)], electron transport flux [C; (ET 0 /RC)] and dissipated energy flux [D; (DI 0 /RC)] from the fluorescence transient exhibited by leaves of rose plants grown under different light spectra [blue (B), red (R), white (W) and red and blue (RB)] under 250 (black bars) and 1500 (grey bars) µmol m −2 s −1 PPFD. Bars represent means ± SD.
Analysis of the parameters that estimate the yields and efficiency of the electron transport chain showed that although the highest values for PI ABS , φ Po and φ Eo were observed in monochromatic R- and B-grown plants under control conditions, their values were decreased more in the HL condition in comparison with their values in W- and RB-grown plants ( Fig. 6A–C ). Exposure to HL led to an increase in φ Do and NPQ [see Supporting Information—Fig. S1 ] in all plants; however, the highest φ Do was observed in RB-grown plants under both control and HL conditions ( Fig. 6D ). An increase in φ RAV was observed in R- and RB-grown plants following exposure to HL ( Fig. 6E ). R- and RB-grown plants had the highest ψ o values under control conditions, while their ψ o values decreased following exposure to HL. Conversely, in W- and RB-grown plants ψ o values were low under control conditions and were increased after exposure to HL ( Fig. 6F ).
Performance index on the absorption basis [A; (PI ABS )], maximum quantum yield of primary photochemistry [B; (φ Po )], quantum yield of electron transport [C; (φ Eo )], quantum yield of energy dissipation [D; (φ Do )], average (from time 0 to t FM ) quantum yield for primary photochemistry [E; (φ PAV )] and the probability that a trapped exciton moves an electron in the electron transport chain beyond QA − [F; (ψ o )] from the fluorescence transient exhibited by leaves of rose plants grown under different light spectra [blue (B), red (R), white (W) and red and blue (RB)] under 250 (black bars) and 1500 (grey bars) µmol m −2 s −1 PPFD. Bars represent means ± SD.
All photosynthetic pigments were significantly influenced by the light spectra ( Fig. 7 ). Chl a [see Supporting Information—Fig. S2A ] , b [see Supporting Information—Fig. S2B ] and total Chl [see Supporting Information—Fig. S2C ] as well as carotenoids ( Fig. 7A ) were significantly decreased by growing rose plants under blue light. In the case of anthocyanin, the interaction between light spectra and light intensity was significant ( P ≤ 0.01). Before exposure to HL, the highest anthocyanin concentration was observed in R- and B-grown plants, while exposure to HL led to an ~50 % reduction in anthocyanin concentration in both R- and B-grown plants; significant differences were not found for anthocyanin concentration before and after HL in the leaves of W- and RB-grown plants ( Fig. 7B ).
Carotenoid (A) and anthocyanin (B) concentrations and MDA (C) content in the leaves of rose plants grown under different light spectra [blue (B), red (R), white (W) and red and blue (RB)] under 250 (black bars) and 1500 (grey bars) µmol m −2 s −1 PPFD. Bars represent means ± SD.
Antioxidant enzymes and oxidative damage
Ascorbate peroxidase and SOD were significantly influenced by light intensity in contrasting ways ( Fig. 8A and B ). Irrespective of the light spectra, APX activity was reduced following exposure to HL, while SOD activity was induced by HL exposure. Catalase activity was reduced by exposure to HL ( Fig. 8C ). The highest CAT activity was observed in RB-grown plants in both control and HL conditions. The H 2 O 2 level was considerably influenced by light intensity ( Fig. 8D ). In comparison with control conditions, the H 2 O 2 level was increased by 10 times following exposure to HL. Malondialdehyde content was also influenced by light intensity ( Fig. 7C ). In comparison with control conditions, the MDA content was doubled after exposure to HL. The highest MDA content was observed in RB-grown plants in both control and HL conditions.
Activity of APX (A), SOD (B) and CAT (C) and H 2 O 2 concentration in the leaves of rose plants grown under different light spectra [blue (B), red (R), white (W) and red and blue (RB)] under 250 (black bars) and 1500 (grey bars) µmol m −2 s −1 PPFD. Bars represent means ± SD.
Carbohydrate concentrations
Soluble carbohydrates were considerably decreased by exposure to HL conditions ( Fig. 9A ). This decrease was greater in R- and B-grown plants (65 % and 79 % decrease, respectively) than in W- and RB-grown plants (60 % and 49 % decrease, respectively). The highest and lowest concentrations of soluble carbohydrates were detected in R- and B-grown plants under control and HL conditions, respectively.
Concentration of soluble carbohydrate (A) and starch (B) in the leaves of rose plants grown under different light spectra [blue (B), red (R), white (W) and red and blue (RB)] under 250 (black bars) and 1500 (grey bars) µmol m −2 s −1 PPFD. Bars represent means ± SD.
Similarly to soluble carbohydrates, HL conditions led to a decrease in starch concentration ( Fig. 9B ). This decrease was, respectively, 28 % and 31 % for R- and B-grown plants and 46 % and 35 % for W- and RB-grown plants. The highest and lowest concentrations of starch were, respectively, detected in R- and W-grown plants under both control and HL conditions.
The reactions involved in photosynthesis in plants are dependent on environmental conditions. Different light attributes including spectrum and intensity are considered to have the most direct effects on photosynthetic reactions ( Chen et al. 2004 ). In our study, Chl fluorescence data were analysed using the OJIP test. Using this test we showed that the PSII system appropriately operated in plants grown under monochromatic R and B spectra. However, when R- and B-grown plants were exposed to HL conditions, their PI ABS and ϕ Eo decreased and DI 0 /RC increased in comparison with their values under control conditions ( Figs 5–7 ). On the other hand, in W- and RB-grown plants ψ o and ET 0 /RC increased following exposure to HL ( Figs 5 and 7 ). These results revealed that growing plants under combination of light spectra enabled the plants to develop a photosynthesis apparatus with lower vulnerability than the photosynthesis apparatus in monochromatically grown plants in response to HL stress. The PI ABS amalgamates the energy fluxes from the early step of the absorption process until the plastoquinone reduction ( Strasser et al. 2000a ). According to previous reports this parameter is very sensitive to different environmental stresses and has been successfully used to measure photosynthetic and plant performance in response to different abiotic stresses including high temperatures ( Martinazzo et al. 2012 ), salinity ( Mathur et al. 2013 ), nutrient deficiency ( Kalaji et al. 2014 ), submergence ( Sarkar and Ray 2016 ), cold ( Rabara et al. 2017 ) and low pH ( Long et al. 2017 ).
In the current experiment, to ensure the same temperature before and after HL, same temperature (27 °C) was also kept inside the growth chambers, which negatively influenced the Chl a fluorescence parameters, resulting in values slightly lower than 0.8 for F v / F m ( Fig. 4 ). The lower PSII photochemical performance (lower PI ABS and φ Eo ) in monochromatic R- and B-grown plants following exposure to HL was due to the higher light energy absorption (ABS/RC), trapping (TR 0 /RC) and dissipated energy flux (DI 0 /RC) and lower ET 0 /ABS per reaction centre ( Fig. 5 ), which consequently resulted in a decreased quantum yield of electron transport (φ Eo ) and maximum quantum yield of primary photochemistry (φ Po ) ( Fig. 6 ). The increase in ABS/RC could attribute to the inactivation of reaction centres and a decrease in active QA reducing centres ( Strasser and Stirbet 1998 ). Under HL conditions, defective QA reducing centres function as a heat sink and protect the plant from HL damage ( Zivcak et al. 2014 ). The higher value of TR 0 /RC resulted in higher inhibition of reoxidation of QA – to QA ( Strasser et al. 2000a ). Higher TR 0 /RC would result in lower electron transport per reaction centre (ET 0 /RC) ( Fig. 5 ), in turn resulting in reduced electron transport per trapping. In fact, a low proportion of absorbed energy is conveyed on the electron transport chain ( Sarkar and Ray 2016 ). Limitation of electron transport beyond PSII will result in QA over-reduction. High light restricts electron transport towards the cytb6f complex and causes QA over-reduction ( Foyer et al. 2012 ). In W- and RB-grown plants, electron transport flux per reaction centre and the probability that a trapped exciton would move an electron in the electron transport chain beyond QA − (ψ o ) increased following HL stress ( Figs 6 and 7 ). This showed that W- and RB-grown plants were more capable of transporting electrons from absorbed photons into the electron transport chain and beyond QA −1 . This confirmed that W- and RB-grown plants were positively regulating the energy level in reaction centres ( Strasser et al. 2004 ) following exposure to HL.
Plants’ adaptation to the prevailing light environment is reflected by a change in anatomy and morphology of leaves, which is known as developmental acclimation ( Vialet-Chabrand et al. 2017 ). Rapid and short-term exposure to HL intensities (as happened in the current study) results in the dynamic acclimation of plants to the light environment ( Lawson et al. 2012 ). Dynamic responses to HL intensities involve metabolic modifications, the down-regulation of the electron transport chain, and changes in stomatal responses and the activation rate of the a ( Tinoco-Ojanguren and Pearcy 1995 ; Lawson et al. 2012 ). In the current study, the focus was on dynamic responses to HL intensities. Therefore, metabolic modifications and changes in performance of the electron transport system were studied in details. Down-regulation of the electron transport system was an obvious response of the photosynthesis system to HL conditions. The antenna in association with PSII is a highly dynamic complex; it collects and transfers the light energy to the PSII reaction centre; however, depending on the physiological needs of the plant, it has the ability to decrease the amount of deliverable light energy to tune the excitation of the reaction centre ( Horton et al. 1996 ). To cope with HL stress, plants dissipate excess energy in the form of heat by a process known as NPQ of chlorophyll fluorescence. Using NPQ, the excitation energy of chlorophyll molecules in the antennae is not passed to the reaction centre and will be dissipated as heat ( Krause and Weis 1991 ). Consistent with this, in the current study, exposure to HL conditions resulted in an increase in NPQ [see Supporting Information—Fig. S1 ] .
Metabolic modifications due to HL were studied by analysing carbohydrate and pigment concentrations at the same time as analysing components of oxidative stress. Soluble carbohydrates and starch concentrations were dramatically decreased after exposure to HL conditions ( Fig. 9 ). Although light is the driving force of photosynthesis for the production of carbohydrates, our result showed that in plants grown under illumination of 250 μmol m −2 s −1 subsequent exposure to HL conditions (1500 μmol m −2 s −1 ) dramatically reduced both soluble and storage carbohydrates ( Fig. 9 ). Positive relationships were found between both soluble carbohydrates and starch with PI ABS ( Fig. 10 ). This indicates that the performance of the PSII operating system plays an important role in soluble carbohydrate and starch production in the rose plant. The greater decrease in soluble carbohydrates in R- and B-grown plants following HL stress is indicative of the higher sensitivity of their PSII operating system to HL stress.
Relationship between performance index on the absorption basis (PI ABS ) and concentration of carbohydrates (soluble carbohydrates and starch) in the leaves of rose plants grown under different light spectra [blue (B), red (R), white (W) and red and blue (RB)] under 250 (C) and 1500 (HL) µmol m −2 s −1 PPFD.
In our study, H 2 O 2 accumulation was detected following HL stress ( Fig. 8 ). The production of ROS following exposure to various environmental stress conditions has previously been reported ( Steffens 2014 ; Pospíšil 2016 ). In photosynthetic electron transport systems, excess light energy beyond photosynthetic capacity induces the production of ROS ( Pospíšil 2016 ). Non-enzymatic and enzymatic scavenging systems are involved in the removal of ROS formed in the electron transport system during exposure to stress conditions ( Ahmad et al. 2010 ). During HL stress, the balance between the scavenging system and ROS formation can be disturbed, leading to oxidative damage to the PSII operating system and its components ( Aro et al. 1993 ). Moreover, in such circumstances, the excess energy absorbed by the PSII complex is not coupled with electron transport, which results in full reduction of the PQ pool and blockage of the electron ( Pospíšil 2016 ). To decrease the reduction of the PQ pool, an electron from Q A •− is transported to an O 2 molecule and generates a superoxide anion. A superoxide anion radical is also generated during the Mehler reaction at the acceptor side of photosystem I. This radical is converted by SOD to H 2 O 2 and subsequently to H 2 O by CAT and APX ( Caverzan et al. 2012 ). We showed that in contrast with a positive correlation between SOD activity and H 2 O 2 production ( Fig. 11 ), CAT and APX negatively regulate H 2 O 2 production. The accumulation of H 2 O 2 following HL stress could be attributed to an increase in SOD activity and decrease in CAT and APX activities ( Fig. 8 ). Photoinactivation of CAT by HL intensities has previously been reported. In accordance with our results, Feierabend and colleagues (1996) showed that CAT is photoinactivated when exposed to B or R lights.
Relationship between H 2 O 2 concentration and activity of antioxidant enzymes (CAT, APX and SOD) in the leaves of rose plants grown under different light spectra [blue (B), red (R), white (W) and red and blue (RB)] under 250 (C) and 1500 (HL) µmol m −2 s −1 PPFD.
In our study, negative relationships were discovered between carbohydrates (soluble carbohydrates and starch) and H 2 O 2 concentration ( Fig. 12 ). In previous studies, it has been reported that sugar accumulation occurs when plants are grown under high illuminations. The relationship between sugars and ROS accumulation is not just a simple positive relationship. Both acceleration of certain ROS-production pathways and deceleration of other ROS-production pathways have been reported due to high levels of soluble carbohydrates ( Couée et al. 2006 ).
Relationship between H 2 O 2 concentration and concentration of carbohydrates (soluble carbohydrates and starch) in the leaves of rose plants grown under different light spectra [blue (B), red (R), white (W) and red and blue (RB)] under 250 (C) and 1500 (HL) µmol m −2 s −1 PPFD.
We found a positive relationship between anthocyanin concentrations and PI ABS under our experimental set-up ( Fig. 13 ). Our data revealed that although plants grown under monochromatic R and B lights had higher anthocyanin concentrations which resulted in better performance of their PSII operating system, their anthocyanin concentrations were dramatically decreased following exposure to HL stress and as a result their PI ABS were also decreased accordingly. In contrast, anthocyanin concentrations in W- and RB-grown plants were not changed before or after HL stress. Therefore, anthocyanin protects the photosynthetic apparatus against HL stress. Apparently, the negative relationship between soluble carbohydrates and ROS accumulation is dependent on the type and duration of the stress condition. For instance, exposure to low temperatures under normal to high irradiances would result in photo-oxidative damage ( Harvaux and Kloppstech 2001 ) and accumulation of soluble carbohydrates. Under such conditions, sugar accumulation can help plants to cope with HL damage ( Ciereszko et al. 2001 ). Carbohydrate accumulation can induce the production of anthocyanin. Chalcone synthase as the key enzyme in anthocyanin biosynthesis can be activated by the signal from soluble carbohydrates ( Koch 1996 ; Deng et al. 2014 ). Therefore, HL stress can induce the accumulation of both anthocyanins and ROS ( Steyn et al. 2002 ; Hatier and Gould 2008 ). In our study, the level of soluble carbohydrates and anthocyanins decreased following exposure to HL to both monochromatic lights, while the levels of these reductions were lower in W- and RB-grown plants. As a consequence, there was lower damage to their PSII operating system ( Figs 10 and 13 ). Anthocyanins can act as light filters and protect the photosynthetic components against photoinhibition induced by HL stress ( Landi et al. 2015 ).
Relationship between performance index on the absorption basis (PI ABS ) and pigments (anthocyanin and carotenoid) in the leaves of rose plants grown under different light spectra [blue (B), red (R), white (W) and red and blue (RB)] under 250 (C) and 1500 (HL) µmol m −2 s −1 PPFD.
The management of a favourable lighting environment for rose production has attracted much attention in the last decade. Currently, many research teams all over the world are focusing on issues related to the best light intensity and spectra for rose plants. To prevent HL intensities shading screens are used for rose production. However, both shading and HL can negatively influence plant growth and photosynthesis. Using polyphasic chlorophyll a fluorescence transients (OJIP test), we showed that growing plants under monochromatic R and B lights resulted in higher sensitivity of their PSII operating system to HL compared to the PSII performance of RB- and W-grown plants. Superoxide dismutase activity increased while CAT and APX activity decreased after exposure to HL. This resulted in the accumulation of H 2 O 2 following HL exposure. Negative relationships were discovered between carbohydrates (soluble carbohydrates and starch) and H 2 O 2 concentration, while carbohydrate concentration positively influenced the performance index of the plants. Our study revealed the protective role of anthocyanin for the PSII operating system. The performance capacity of the photosynthetic systems in monochromatic R- and B-grown plants declined due to a dramatic reduction in anthocyanin concentration following HL exposure, whereas W- and RB-grown plants had similar anthocyanin levels before and after HL exposure. This suggests that anthocyanin protected their photosynthetic apparatus against HL stress.
We would like to thank Iran National Science Foundation (INSF) (grant number 96006991) and University of Tehran for their supports.
S.A. made substantial contributions to conception and design, also performed statistical analysis, drafted the manuscript and critically revised the final version. L.B. carried out the experiments, collected and critically analysed the scientific literatures and help in the writing of the manuscript. M.A. took part in designing and planning the experiments, preparation of material for the experiment and contributed to scientific discussion of the obtained results. M.S. helped in preparation of material for the experiment and took part in designing, planning and performing of experiments. T.L. contributed to conception and design of experiment, preparation of materials, scientific discussion and critical revision of the final manuscript. O.L. contributed to design of experiment and critical revision of the final manuscript.
None declared.
The following additional information is available in the online version of this article—
Table S1. Analysis of variance ( F -values) for assessed parameters for rose plants grown under different light spectrums and then exposed to high light intensity (1500 μmol m −2 s −1 ).
Figure S1. Non-photochemical quenching (NPQ) derived from chlorophyll fluorescence parameters in the leaves of rose plants grown at different light spectrums [blue (B), red (R), white (W) and red and blue (RB)] under 250 (black bars) and 1500 (grey bars) µmol m −2 s −1 photosynthetic photon flux density (PPFD). Bars represent means ± SD.
Figure S2. Chlorophyll a (A), chlorophyll b (B) and total chlorophyll (C) concentrations in the leaves of rose plants grown at different light spectrums [blue (B), red (R), white (W) and red and blue (RB)] under 250 (black bars) and 1500 (grey bars) µmol m −2 s −1 photosynthetic photon flux density (PPFD). Bars represent means ± SD.
We thank Z. Soltani, K. Pirasteh and S. Moemeni for their assistance in performing the experiments and measurements, Dr S. Kalhor for critical revision of the manuscript and Iran Grow Light Company for providing LEDs and projectors.
Ahmad P , Jaleel CA , Salem MA , Nabi G , Sharma S . 2010 . Roles of enzymatic and nonenzymatic antioxidants in plants during abiotic stress . Critical Reviews in Biotechnology 30 : 161 – 175 .
Google Scholar
Aliniaeifard S , Malcolm Matamoros P , van Meeteren U . 2014 . Stomatal malfunctioning under low VPD conditions: induced by alterations in stomatal morphology and leaf anatomy or in the ABA signaling ? Physiologia Plantarum 152 : 688 – 699 .
Aliniaeifard S , van Meeteren U . 2014 . Natural variation in stomatal response to closing stimuli among Arabidopsis thaliana accessions after exposure to low VPD as a tool to recognize the mechanism of disturbed stomatal functioning . Journal of Experimental Botany 65 : 6529 – 6542 .
Arnon DI . 1949 . Copper enzymes in isolated chloroplasts. Polyphenoloxidase in Beta vulgaris . Plant Physiology 24 : 1 – 15 .
Aro EM , Virgin I , Andersson B . 1993 . Photoinhibition of photosystem II. Inactivation, protein damage and turnover . Biochimica et Biophysica Acta (BBA)-Bioenergetics 1143 : 113 – 134 .
Breive K , Tamulaitis G , Duchovskis P , Bliznikas Z , Ulinskaite R , Brazaityte A , Novickovas A , Zukauskas A . 2005 . High-power light-emitting diode based facility for plant cultivation . Journal of Physics D: Applied Physics 38 : 3182 – 3187 .
Brestic M , Zivcak M . 2013 . PSII fluorescence techniques for measurement of drought and high temperature stress signal in crop plants: protocols and applications . In Molecular stress physiology of plants . India : Springer , 87 – 131 .
Google Preview
Carvalho SD , Schwieterman ML , Abrahan CE , Colquhoun TA , Folta KM . 2016 . Light quality dependent changes in morphology, antioxidant capacity, and volatile production in sweet basil ( Ocimum basilicum ) . Frontiers in Plant Science 7 : 1328 .
Caverzan A , Passaia G , Rosa SB , Ribeiro CW , Lazzarotto F , Margis-Pinheiro M . 2012 . Plant responses to stresses: role of ascorbate peroxidase in the antioxidant protection . Genetics and Molecular Biology 35 : 1011 – 1019 .
Chen M , Chory J , Fankhauser C . 2004 . Light signal transduction in higher plants . Annual Review of Genetics 38 : 87 – 117 .
Ciereszko I , Johansson H , Kleczkowski LA . 2001 . Sucrose and light regulation of a cold-inducible UDP-glucose pyrophosphorylase gene via a hexokinase-independent and abscisic acid-insensitive pathway in Arabidopsis . The Biochemical Journal 354 : 67 – 72 .
Couée I , Sulmon C , Gouesbet G , El Amrani A . 2006 . Involvement of soluble sugars in reactive oxygen species balance and responses to oxidative stress in plants . Journal of Experimental Botany 57 : 449 – 459 .
Dat J , Vandenabeele S , Vranová E , Van Montagu M , Inzé D , Van Breusegem F . 2000 . Dual action of the active oxygen species during plant stress responses . Cellular and Molecular Life Sciences: CMLS 57 : 779 – 795 .
Deng X , Bashandy H , Ainasoja M , Kontturi J , Pietiäinen M , Laitinen RA , Albert VA , Valkonen JP , Elomaa P , Teeri TH . 2014 . Functional diversification of duplicated chalcone synthase genes in anthocyanin biosynthesis of Gerbera hybrida . The New Phytologist 201 : 1469 – 1483 .
Dhindsa R , Plumb-Dhindsa P , Thorpe T . 1981 . Leaf senescence: correlated with increased levels of membrane permeability and lipid peroxidation, and decreased levels of superoxide dismutase and catalase . Journal of Experimental Botany 32 : 93 – 101 .
Díaz-Vivancos P , Clemente-Moreno MJ , Rubio M , Olmos E , García JA , Martínez-Gómez P , Hernández JA . 2008 . Alteration in the chloroplastic metabolism leads to ROS accumulation in pea plants in response to plum pox virus . Journal of Experimental Botany 59 : 2147 – 2160 .
Fanourakis D , Pieruschka R , Savvides A , Macnish AJ , Sarlikioti V , Woltering EJ . 2013 . Sources of vase life variation in cut roses: a review . Postharvest Biology and Technology 78 : 1 – 15 .
Feierabend J , Streb P , Schmidt M , Dehne S , Shang W . 1996 . Expression of catalase and its relation to light stress and stress tolerance . In: Physical stresses in plants . Berlin : Springer , 223 – 234 .
Foyer CH , Neukermans J , Queval G , Noctor G , Harbinson J . 2012 . Photosynthetic control of electron transport and the regulation of gene expression . Journal of Experimental Botany 63 : 1637 – 1661 .
García Victoria N , Kempkes F , Van Weel P , Stanghellini C , Dueck T , Bruins M . 2011 . Effect of a diffuse glass greenhouse cover on rose production and quality . In: International Symposium on Advanced Technologies and Management Towards Sustainable Greenhouse Ecosystems. Greensys2011 952 , 241 – 248 . doi:10.17660/ActaHortic.2012.952.29
Genty B , Briantais J-M , Baker NR . 1989 . The relationship between the quantum yield of photosynthetic electron transport and quenching of chlorophyll fluorescence . Biochimica et Biophysica Acta (BBA)-General Subjects 990 : 87 – 92 .
Gu J , Zhou Z , Li Z , Chen Y , Wang Z , Zhang H , Yang J . 2017 . Photosynthetic properties and potentials for improvement of photosynthesis in pale green leaf rice under high light conditions . Frontiers in Plant Science 8 . doi:10.3389/fpls.2017.01082
Harvaux M , Kloppstech K . 2001 . The protective functions of carotenoid and flavonoid pigments against excess visible radiation at chilling temperature investigated in arabidopsis npq and tt mutants . Planta 213 : 953 – 966 .
Hatier J-HB , Gould KS . 2008 . Anthocyanin function in vegetative organs . In Anthocyanins . Springer , 1 – 19 . doi:10.1016/0003-9861(68)90654-1
Heath RL , Packer L . 1968 . Photoperoxidation in isolated chloroplasts. I. Kinetics and stoichiometry of fatty acid peroxidation . Archives of Biochemistry and Biophysics 125 : 189 – 198 .
Horton P , Ruban AV , Walters RG . 1996 . Regulation of light harvesting in green plants . Annual Review of Plant Physiology and Plant Molecular Biology 47 : 655 – 684 .
Kalaji MH , Goltsev VN , Zivcak M , Brestic M . 2017 . Chlorophyll fluorescence: understanding crop performance—basics and applications . Boca Raton : CRC Press .
Kalaji HM , Jajoo A , Oukarroum A , Brestic M , Zivcak M , Samborska IA , Cetner MD , Łukasik I , Goltsev V , Ladle RJ . 2016 . Chlorophyll a fluorescence as a tool to monitor physiological status of plants under abiotic stress conditions . Acta Physiologiae Plantarum 38 : 102
Kalaji HM , Oukarroum A , Alexandrov V , Kouzmanova M , Brestic M , Zivcak M , Samborska IA , Cetner MD , Allakhverdiev SI , Goltsev V . 2014 . Identification of nutrient deficiency in maize and tomato plants by in vivo chlorophyll a fluorescence measurements . Plant Physiology and Biochemistry 81 : 16 – 25 .
Kim SJ , Hahn EJ , Heo JW , Paek KY . 2004 . Effects of LEDs on net photosynthetic rate, growth and leaf stomata of chrysanthemum plantlets in vitro . Scientia Horticulturae 101 : 143 – 151
Koch KE . 1996 . Carbohydrate-modulated gene expression in plants . Annual Review of Plant Physiology and Plant Molecular Biology 47 : 509 – 540 .
Kozai T . 2016 . LED lighting for urban agriculture . In: Kozai T , ed. LED lighting for urban agriculture . Singapore : Springer .
Kozai T , Niu G , Takagaki M . 2015 . Plant factory: an indoor vertical farming system for efficient quality food production . USA : Academic Press .
Krause G , Weis E . 1991 . Chlorophyll fluorescence and photosynthesis: the basics . Annual Review of Plant Physiology and Plant Molecular Biology 42 : 313 – 349 .
Landi M , Tattini M , Gould KS . 2015 . Multiple functional roles of anthocyanins in plant-environment interactions . Environmental and Experimental Botany 119 : 4 – 17
Lawson T , Kramer DM , Raines CA . 2012 . Improving yield by exploiting mechanisms underlying natural variation of photosynthesis . Current Opinion in Biotechnology 23 : 215 – 220 .
Li Q , Kubota C . 2009 . Effects of supplemental light quality on growth and phytochemicals of baby leaf lettuce . Environmental and Experimental Botany 67 : 59 – 64
Long A , Zhang J , Yang LT , Ye X , Lai NW , Tan LL , Lin D , Chen LS . 2017 . Effects of low pH on photosynthesis, related physiological parameters, and nutrient profiles of citrus . Frontiers in Plant Science 8 : 185 .
Martinazzo EG , Ramm A , Bacarin MA . 2012 . The chlorophyll a fluorescence as an indicator of the temperature stress in the leaves of Prunus persica . Brazilian Journal of Plant Physiology 24 : 237 – 246
Mathur S , Mehta P , Jajoo A . 2013 . Effects of dual stress (high salt and high temperature) on the photochemical efficiency of wheat leaves ( Triticum aestivum ) . Physiology and Molecular Biology of Plants 19 : 179 – 188 .
McCready R , Guggolz J , Silviera V , Owens H . 1950 . Determination of starch and amylose in vegetables . Analytical Chemistry 22 : 1156 – 1158 .
Miyake C , Amako K , Shiraishi N , Sugimoto T . 2009 . Acclimation of tobacco leaves to high light intensity drives the plastoquinone oxidation system–relationship among the fraction of open PSII centers, non-photochemical quenching of chl fluorescence and the maximum quantum yield of PSII in the dark . Plant & Cell Physiology 50 : 730 – 743 .
Morrow RC . 2008 . LED lighting in horticulture . Hortscience 43 : 1947 – 1950 .
Müller P , Li XP , Niyogi KK . 2001 . Non-photochemical quenching. A response to excess light energy . Plant Physiology 125 : 1558 – 1566 .
Nakano Y , Asada K . 1981 . Hydrogen peroxide is scavenged by ascorbate-specific peroxidase in spinach chloroplasts . Plant and Cell Physiology 22 : 867 – 880 .
Panda D , Sharma S , Sarkar R . 2008 . Fast chlorophyll fluorescence transients as selection tools for submergence tolerance in rice ( Oryza sativa ) . Indian Journal of Agricultural Sciences 78 : 933 .
Patterson BD , MacRae EA , Ferguson IB . 1984 . Estimation of hydrogen peroxide in plant extracts using titanium (IV) . Analytical Biochemistry 139 : 487 – 492 .
Petrella DP , Metzger JD , Blakeslee JJ , Nangle EJ , Gardner DS . 2016 . Anthocyanin production using rough bluegrass treated with high-intensity light . Hortscience 51 : 1111 – 1120 .
Pospíšil P . 2016 . Production of reactive oxygen species by photosystem II as a response to light and temperature stress . Frontiers in Plant Science 7 . doi:10.3389/fpls.2016.01950
Rabara RC , Behrman G , Timbol T , Rushton PJ . 2017 . Effect of spectral quality of monochromatic LED lights on the growth of artichoke seedlings . Frontiers in Plant Science 8 : 190 .
Rapacz M , Sasal M , Kalaji HM , Kościelniak J . 2015 . Is the OJIP test a reliable indicator of winter hardiness and freezing tolerance of common wheat and triticale under variable winter environments ? PLoS One 10 : e0134820 .
Roach T , Krieger-Liszkay A . 2014 . Regulation of photosynthetic electron transport and photoinhibition . Current Protein & Peptide Science 15 : 351 – 362 .
Sæbø A , Krekling T , Appelgren M . 1995 . Light quality affects photosynthesis and leaf anatomy of birch plantlets in vitro . Plant Cell, Tissue and Organ Culture 41 : 177 – 185 .
Sarkar R , Ray A . 2016 . Submergence-tolerant rice withstands complete submergence even in saline water: probing through chlorophyll a fluorescence induction OJIP transients . Photosynthetica 54 : 275 – 287 .
Sharma PK , Hall DO . 1993 . The role of carotenoids in protection against photoinhibition . In: Abrol YP , Mohanty P. , Govindjee , eds. Photosynthesis: photoreactions to plant productivity . Dordrecht : 469 – 478 .
Smith H . 1982 . Light quality, photoperception, and plant strategy . Annual Review of Plant Physiology 33 : 481 – 518 .
Steffens B . 2014 . The role of ethylene and ROS in salinity, heavy metal, and flooding responses in rice . Frontiers in Plant Science 5 : 685 .
Steyn W , Wand S , Holcroft D , Jacobs G . 2002 . Anthocyanins in vegetative tissues: a proposed unified function in photoprotection . New Phytologist 155 : 349 – 361 .
Strasser BJ . 1995 . Measuring fast fluorescence transients to address environmental questions: the JIP test . In: Mathis , Paul , eds. Photosynthesis: from light to biosphere . Dordrecht, the Netherlands : Kluwer Academic Publishers , 977 – 980 .
Strasser RJ , Srivastava A , Tsimilli-Michael M . 2000a . The fluorescence transient as a tool to characterize and screen photosynthetic samples . In: Mohammad Yunus , Uday Pathre , Prasanna Mohanty , eds. Probing photosynthesis: mechanisms, regulation and adaptation . London and New York : CRC press , 445 – 483 .
Strasser RJ , Srivastava A , Tsimilli-Michael MG . 2000b . The fluorescence transient as a tool to characterize and screen photosynthetic samples . In: Mohammad Yunus , Uday Pathre , Prasanna Mohanty , eds. Probing photosynthesis: mechanisms, regulation and adaptation . London and New York : CRC press , 445 – 483 .
Strasser RJ , Stirbet AD . 1998 . Heterogeneity of photosystem II probed by the numerically simulated chlorophyll a fluorescence rise (O–J–I–P) . Mathematics and Computers in Simulation 48 : 3 – 9 .
Strasser RJ , Tsimilli-Michael M , Srivastava A . 2004 . Analysis of the chlorophyll a fluorescence transient . In: George Christos Papageorgiou , Govindjee , eds. Chlorophyll a fluorescence . Dordrecht : Springer , 321 – 362 .
Taiz L , Zeiger E . 2002 . Plant physiology . USA : Sinauer Associates .
Tennessen DJ , Singsaas EL , Sharkey TD . 1994 . Light-emitting diodes as a light source for photosynthesis research . Photosynthesis Research 39 : 85 – 92 .
Tinoco-Ojanguren C , Pearcy R . 1995 . A comparison of light quality and quantity effects on the growth and steady-state and dynamic photosynthetic characteristics of three tropical tree species . Functional Ecology 9 : 222 – 230 .
van Doorn WG . 2012 . Water relations of cut flowers: an update . Horticultural Reviews 40 : 55 – 106 .
Vass I . 2012 . Molecular mechanisms of photodamage in the photosystem II complex . Biochimica et Biophysica Acta (BBA)-Bioenergetics 1817 : 209 – 217 .
Vialet-Chabrand S , Matthews JS , Simkin AJ , Raines CA , Lawson T . 2017 . Importance of fluctuations in light on plant photosynthetic acclimation . Plant Physiology 173 : 2163 – 2179 .
Vranová E , Inzé D , Van Breusegem F . 2002 . Signal transduction during oxidative stress . Journal of Experimental Botany 53 : 1227 – 1236 .
Wagner GJ . 1979 . Content and vacuole/extravacuole distribution of neutral sugars, free amino acids, and anthocyanin in protoplasts . Plant Physiology 64 : 88 – 93 .
Whitelam GC , Halliday KJ . 2008 . Annual plant reviews, light and plant development . UK : John Wiley & Sons .
Zhao X , Chen T , Feng B , Zhang C , Peng S , Zhang X , Fu G , Tao L . 2017 . Non-photochemical quenching plays a key role in light acclimation of rice plants differing in leaf color . Frontiers in Plant Science 7 . doi:10.3389/fpls.2016.01968
Zivcak M , Brestic M , Kalaji HM , Govindjee . 2014 . Photosynthetic responses of sun- and shade-grown barley leaves to high light: is the lower PSII connectivity in shade leaves associated with protection against excess of light ? Photosynthesis Research 119 : 339 – 354 .
Supplementary data
Month: | Total Views: |
---|---|
September 2018 | 100 |
October 2018 | 248 |
November 2018 | 257 |
December 2018 | 190 |
January 2019 | 264 |
February 2019 | 582 |
March 2019 | 917 |
April 2019 | 1,083 |
May 2019 | 1,665 |
June 2019 | 1,602 |
July 2019 | 950 |
August 2019 | 1,013 |
September 2019 | 1,697 |
October 2019 | 2,803 |
November 2019 | 2,666 |
December 2019 | 1,920 |
January 2020 | 1,818 |
February 2020 | 2,986 |
March 2020 | 1,892 |
April 2020 | 1,363 |
May 2020 | 1,327 |
June 2020 | 1,883 |
July 2020 | 1,298 |
August 2020 | 1,579 |
September 2020 | 2,590 |
October 2020 | 4,222 |
November 2020 | 4,082 |
December 2020 | 3,029 |
January 2021 | 2,946 |
February 2021 | 2,947 |
March 2021 | 3,573 |
April 2021 | 3,062 |
May 2021 | 2,295 |
June 2021 | 1,577 |
July 2021 | 1,097 |
August 2021 | 1,213 |
September 2021 | 1,991 |
October 2021 | 2,621 |
November 2021 | 2,559 |
December 2021 | 1,652 |
January 2022 | 1,545 |
February 2022 | 1,898 |
March 2022 | 2,065 |
April 2022 | 1,738 |
May 2022 | 1,520 |
June 2022 | 822 |
July 2022 | 589 |
August 2022 | 698 |
September 2022 | 1,655 |
October 2022 | 1,986 |
November 2022 | 1,416 |
December 2022 | 899 |
January 2023 | 686 |
February 2023 | 779 |
March 2023 | 805 |
April 2023 | 572 |
May 2023 | 662 |
June 2023 | 349 |
July 2023 | 311 |
August 2023 | 256 |
September 2023 | 592 |
October 2023 | 699 |
November 2023 | 618 |
December 2023 | 420 |
January 2024 | 436 |
February 2024 | 424 |
March 2024 | 451 |
April 2024 | 326 |
May 2024 | 351 |
June 2024 | 184 |
July 2024 | 211 |
August 2024 | 450 |
September 2024 | 473 |
October 2024 | 268 |
Email alerts
Citing articles via.
- X (formerly Twitter)
Affiliations
- Online ISSN 2041-2851
- Copyright © 2024 Annals of Botany Company
- About Oxford Academic
- Publish journals with us
- University press partners
- What we publish
- New features
- Open access
- Institutional account management
- Rights and permissions
- Get help with access
- Accessibility
- Advertising
- Media enquiries
- Oxford University Press
- Oxford Languages
- University of Oxford
Oxford University Press is a department of the University of Oxford. It furthers the University's objective of excellence in research, scholarship, and education by publishing worldwide
- Copyright © 2024 Oxford University Press
- Cookie settings
- Cookie policy
- Privacy policy
- Legal notice
This Feature Is Available To Subscribers Only
Sign In or Create an Account
This PDF is available to Subscribers Only
For full access to this pdf, sign in to an existing account, or purchase an annual subscription.
Advertisement
The role of photosynthesis related pigments in light harvesting, photoprotection and enhancement of photosynthetic yield in planta
- Published: 22 January 2022
- Volume 152 , pages 23–42, ( 2022 )
Cite this article
- Andrew J. Simkin 1 na1 ,
- Leepica Kapoor 2 na1 ,
- C. George Priya Doss 2 ,
- Tanja A. Hofmann 4 ,
- Tracy Lawson 3 &
- Siva Ramamoorthy 2
7515 Accesses
5 Altmetric
Explore all metrics
Photosynthetic pigments are an integral and vital part of all photosynthetic machinery and are present in different types and abundances throughout the photosynthetic apparatus. Chlorophyll, carotenoids and phycobilins are the prime photosynthetic pigments which facilitate efficient light absorption in plants, algae, and cyanobacteria. The chlorophyll family plays a vital role in light harvesting by absorbing light at different wavelengths and allowing photosynthetic organisms to adapt to different environments, either in the long-term or during transient changes in light. Carotenoids play diverse roles in photosynthesis, including light capture and as crucial antioxidants to reduce photodamage and photoinhibition. In the marine habitat, phycobilins capture a wide spectrum of light and have allowed cyanobacteria and red algae to colonise deep waters where other frequencies of light are attenuated by the water column. In this review, we discuss the potential strategies that photosynthetic pigments provide, coupled with development of molecular biological techniques, to improve crop yields through enhanced light harvesting, increased photoprotection and improved photosynthetic efficiency.
This is a preview of subscription content, log in via an institution to check access.
Access this article
Subscribe and save.
- Get 10 units per month
- Download Article/Chapter or eBook
- 1 Unit = 1 Article or 1 Chapter
- Cancel anytime
Price includes VAT (Russian Federation)
Instant access to the full article PDF.
Rent this article via DeepDyve
Institutional subscriptions
Similar content being viewed by others
Molecular Mechanism of Photosynthesis Driven by Red-Shifted Chlorophylls
Far-red light photoadaptations in aquatic cyanobacteria.
Recent Advances in the Photosynthesis of Cyanobacteria and Eukaryotic Algae
Data availability.
All data supporting the findings of this study are available within the paper.
Code availability
Not applicable.
Allakhverdiev S, Kreslavski V, Zharmukhamedov S, Voloshin R, Korol’kova D, Tomo T, Shen J-R (2016) Chlorophylls d and f and their role in primary photosynthetic processes of cyanobacteria. Biochem Mosc 81(3):201–212
Article CAS Google Scholar
Allen JF, Forsberg J (2001) Molecular recognition in thylakoid structure and function. Trends Plant Sci 6(7):317–326
Article CAS PubMed Google Scholar
Alotaibi SS, Sparks CA, Parry MA, Simkin AJ, Raines CA (2018) Identification of leaf promoters for use in transgenic wheat. Plants 7(2):27
Article PubMed Central CAS Google Scholar
Andreoni A, Lin S, Liu H, Blankenship RE, Yan H, Woodbury NW (2017) Orange carotenoid protein as a control element in an antenna system based on a DNA nanostructure. Nano Lett 17(2):1174–1180
Ariizumi T, Kishimoto S, Kakami R, Maoka T, Hirakawa H, Suzuki Y, Ozeki Y, Shirasawa K, Bernillon S, Okabe Y, Moing A, Asamizu E, Rothan C, Ohmiya A, Ezura H (2014) Identification of the carotenoid modifying gene PALE YELLOW PETAL 1 as an essential factor in xanthophyll esterification and yellow flower pigmentation in tomato (Solanum lycopersicum). Plant J 79(3):453–465. https://doi.org/10.1111/tpj.12570
Armbruster U, Galvis VC, Kunz H-H, Strand DD (2017) The regulation of the chloroplast proton motive force plays a key role for photosynthesis in fluctuating light. Curr Opin Plant Biol 37:56–62
Auldridge ME, Block A, Vogel JT, Dabney-Smith C, Mila I, Bouzayen M, Magallanes-Lundback M, DellaPenna D, McCarty DR, Klee HJ (2006a) Characterization of three members of the Arabidopsis carotenoid cleavage dioxygenase family demonstrates the divergent roles of this multifunctional enzyme family. Plant J 45(6):982–993. https://doi.org/10.1111/j.1365-313X.2006.02666.x
Auldridge ME, McCarty DR, Klee HJ (2006b) Plant carotenoid cleavage oxygenases and their apocarotenoid products. Curr Opin Plant Biol 9(3):315–321
Ayumi Tanaka A (1998) Chlorophyll a oxygenase (CAO) is involved in chlorophyll b formation from chlorophyll a. Proc Natl Acad Sci USA 95:12719–12723
Article PubMed Central Google Scholar
Baker NR (2008) Chlorophyll fluorescence: a probe of photosynthesis in vivo. Annu Rev Plant Biol 59:89–113
Ballottari M, Truong TB, De Re E, Erickson E, Stella GR, Fleming GR, Bassi R, Niyogi KK (2016) Identification of pH-sensing sites in the light harvesting complex stress-related 3 protein essential for triggering non-photochemical quenching in Chlamydomonas reinhardtii . J Biol Chem 291(14):7334–7346
Article CAS PubMed PubMed Central Google Scholar
Bamsey M, Berinstain A, Graham T, Neron P, Giroux R, Braham S, Ferl R, Paul A-L, Dixon M (2009) Developing strategies for automated remote plant production systems: environmental control and monitoring of the Arthur Clarke Mars Greenhouse in the Canadian High Arctic. Adv Space Res 44(12):1367–1381
Bamsey MT, Zabel P, Zeidler C, Gyimesi D, Schubert D, Kohlberg E, Mengedoht D, Rae J, Graham T (2015) Review of Antarctic greenhouses and plant production facilities: A historical account of food plants on the ice. In: 45th international conference on environmental systems
Bassham JA, Calvin M (1960) The path of carbon in photosynthesis. In: Die CO2-assimilation/the assimilation of carbon dioxide. Springer, New York, pp 884–922
Beale S (1999) Enzymes of chlorophyll biosynthesis. Photosynth Res 60:43–73
Beckmann J, Lehr F, Finazzi G, Hankamer B, Posten C, Wobbe L, Kruse O (2009) Improvement of light to biomass conversion by de-regulation of light-harvesting protein translation in Chlamydomonas reinhardtii. J Biotechnol 142(1):70–77
Biel K, Fomina I (2015) Benson-Bassham-Calvin cycle contribution to the organic life on our planet. Photosynthetica 53(2):161–167. https://doi.org/10.1007/s11099-015-0112-7
Blankenship R (2002) Molecular mechanisms of photosynthesis. Blackwell Science, Maiden
Book Google Scholar
Blankenship RE (2010) Early evolution of photosynthesis. Plant Physiol 154(2):434–438. https://doi.org/10.1104/pp.110.161687
Blankenship RE (2014) Molecular mechanisms of photosynthesis. Wiley, New York
Google Scholar
Bonente G, Ballottari M, Truong TB, Morosinotto T, Ahn TK, Fleming GR, Niyogi KK, Bassi R (2011) Analysis of LhcSR3, a protein essential for feedback de-excitation in the green alga Chlamydomonas reinhardtii. PLoS Biol 9(1):e1000577
Booker J, Auldridge M, Wills S, McCarty D, Klee H, Leyser O (2004) MAX3/CCD7 is a carotenoid cleavage dioxygenase required for the synthesis of a novel plant signaling molecule. Curr Biol 14(14):1232–1238. https://doi.org/10.1016/j.cub.2004.06.061
Bouvier F, Dogbo O, Camara B (2003a) Biosynthesis of the food and cosmetic plant pigment bixin (annatto). Science 300(5628):2089–2091
Bouvier F, Suire C, Mutterer J, Camara B (2003b) Oxidative remodeling of chromoplast carotenoids: identification of the carotenoid dioxygenase CsCCD and CsZCD genes involved in Crocus secondary metabolite biogenesis. Plant Cell 15(1):47–62
Budzikiewicz H, Taraz K (1971) Chlorophyll c. Tetrahedron 27(7):1447–1460. https://doi.org/10.1016/S0040-4020(01)98010-X
Bungard RA, Ruban AV, Hibberd JM, Press MC, Horton P, Scholes JD (1999) Unusual carotenoid composition and a new type of xanthophyll cycle in plants. Proc Natl Acad Sci 96(3):1135–1139. https://doi.org/10.1073/pnas.96.3.1135
Calderini DF, Dreccer MF, Slafer GA (1995) Genetic improvement in wheat yield and associated traits. A re-examination of previous results and the latest trends. Plant Breeding 114:108–112
Article Google Scholar
Calvin M, Benson AA (1948) The path of carbon in photosynthesis. Science 107(2784):476–480. https://doi.org/10.1126/science.107.2784.476
Cazzaniga S, Bressan M, Carbonera D, Agostini A, Dall’Osto L (2016) Differential roles of carotenes and xanthophylls in photosystem I photoprotection. Biochemistry 55(26):3636–3649. https://doi.org/10.1021/acs.biochem.6b00425
Chakdar H, Pabbi S (2016) Cyanobacterial phycobilins: production, purification, and regulation. Frontier discoveries and innovations in interdisciplinary microbiology. Springer, New York, pp 45–69
Chapter Google Scholar
Chalker-Scott L (1999) Environmental significance of anthocyanins in plant stress responses. Photochem Photobiol 70(1):1–9
Chen M, Blankenship RE (2011) Expanding the solar spectrum used by photosynthesis. Trends Plant Sci 16(8):427–431
Chen M, Schliep M, Willows RD, Cai ZL, Neilan BA, Scheer H (2010) A red-shifted chlorophyll. Science 329(5997):1318–1319. https://doi.org/10.1126/science.1191127
Chida H, Nakazawa A, Akazaki H, Hirano T, Suruga K, Ogawa M, Satoh T, Kadokura K, Yamada S, Hakamata W, Isobe K, Ito T, Ishii R, Nishio T, Sonoike K, Oku T (2007) Expression of the algal cytochrome c 6 gene in Arabidopsis enhances photosynthesis and growth. Plant Cell Physiol 48(7):948–957. https://doi.org/10.1093/pcp/pcm064
Cho F (1970) Low-temperature (4–77 degrees K) spectroscopy of Chlorella: temperature dependence of energy transfer efficiency. Biochim Biophys Acta 216(1):139–150. https://doi.org/10.1016/0005-2728(70)90166-0
Christie JM (2007) Phototropin blue-light receptors. Annu Rev Plant Biol 58:21–45
Correa-Galvis V, Redekop P, Guan K, Griess A, Truong TB, Wakao S, Niyogi KK, Jahns P (2016) Photosystem II subunit PsbS is involved in the induction of LHCSR protein-dependent energy dissipation in Chlamydomonas reinhardtii . J Biol Chem 291(33):17478–17487. https://doi.org/10.1074/jbc.M116.737312
Croce R (2012) Chlorophyll-binding proteins of higher plants and cyanobacteria. In: Eaton-Rye JJ, Tripathy BC, Sharkey TD (eds) Photosynthesis: plastid biology, energy conversion and carbon assimilation. Springer, Dordrecht, pp 127–149
Croce R, Van Amerongen H (2014) Natural strategies for photosynthetic light harvesting. Nat Chem Biol 10(7):492
Czarnecki O, Grimm B (2012) Post-translational control of tetrapyrrole biosynthesis in plants, algae, and cyanobacteria. J Exp Bot 63(4):1675–1687
Dall’Osto L, Lico C, Alric J, Giuliano G, Havaux M, Bassi R (2006) Lutein is needed for efficient chlorophyll triplet quenching in the major LHCII antenna complex of higher plants and effective photoprotection in vivo under strong light. BMC Plant Biol 6(1):32
Article PubMed PubMed Central CAS Google Scholar
Dall’Osto L, Fiore A, Cazzaniga S, Giuliano G, Bassi R (2007) Different roles of alpha- and beta-branch xanthophylls in photosystem assembly and photoprotection. J Biol Chem 282(48):35056–35068. https://doi.org/10.1074/jbc.M704729200
Dall’Osto L, Cazzaniga S, Havaux M, Bassi R (2010) Enhanced photoprotection by protein-bound vs free xanthophyll pools: a comparative analysis of chlorophyll b and xanthophyll biosynthesis mutants. Mol Plant 3(3):576–593
de Bianchi S, Ballottari M, Dall’osto L, Bassi R (2010) Regulation of plant light harvesting by thermal dissipation of excess energy. Biochem Soc Trans 38(2):651–660. https://doi.org/10.1042/BST0380651
Demmig-Adams B (1998) Survey of thermal energy dissipation and pigment composition in sun and shade leaves. Plant Cell Physiol 39(5):474–482
Demmig-Adams B, Adams WW III (1992) Carotenoid composition in sun and shade leaves of plants with different life forms. Plant Cell Environ 15(4):411–419. https://doi.org/10.1111/j.1365-3040.1992.tb00991.x
Demmig-Adams B, Adams WW III (1996) The role of xanthophyll cycle carotenoids in the protection of photosynthesis. Trends Plant Sci 1(1):21–26
Deruere J, Romer S, d’Harlingue A, Backhaus RA, Kuntz M, Camara B (1994) Fibril assembly and carotenoid overaccumulation in chromoplasts: a model for supramolecular lipoprotein structures. Plant Cell 6(1):119–133. https://doi.org/10.1105/tpc.6.1.119
Domonkos I, Kis M, Gombos Z, Ughy B (2013) Carotenoids, versatile components of oxygenic photosynthesis. Prog Lipid Res 52(4):539–561
Dougherty RC, Strain HH, Svec WA, Uphaus RA, Katz JJ (1970) Structure, properties, and distribution of chlorophyll c. J Am Chem Soc 92(9):2826–2833. https://doi.org/10.1021/ja00712a037
Driever SM, Simkin AJ, Alotaibi S, Fisk SJ, Madgwick PJ, Sparks CA, Jones HD, Lawson T, Parry MAJ, Raines CA (2017) Increased SBPase activity improves photosynthesis and grain yield in wheat grown in greenhouse conditions. Philos Trans R Soc B 372:1730
Engler C, Kandzia R, Marillonnet S (2008) A one pot, one step, precision cloning method with high throughput capability. PLoS ONE 3(11):e3647
Esteban R, Becerril JM, García-Plazaola JI (2009a) Lutein epoxide cycle, more than just a forest tale. Plant Signal Behav 4(4):342–344. https://doi.org/10.4161/psb.4.4.8197
Esteban R, Olano JM, Castresana J, Fernández-Marín B, Hernández A, Becerril JM, García-Plazaola JI (2009b) Distribution and evolutionary trends of photoprotective isoprenoids (xanthophylls and tocopherols) within the plant kingdom. Physiol Plant 135(4):379–389. https://doi.org/10.1111/j.1399-3054.2008.01196.x
Esteban R, Barrutia O, Artetxe U, Fernández-Marín B, Hernández A, García-Plazaola JI (2015) Internal and external factors affecting photosynthetic pigment composition in plants: a meta-analytical approach. New Phytol 206(1):268–280
Evans TL, Fischer RA (1999) Yield potential: its definition, measurement, and significance. Crop Sci 39:1544–1551
Exposito-Rodriguez M, Laissue PP, López-Calcagno PE, Mullineaux PM, Raines CA, Simkin AJ (2017) Development of pGEMINI, a plant gateway destination vector allowing the simultaneous integration of two cDNA via a single LR-clonase reaction. Plants 6(4):55
Fan Y, Chen J, Cheng Y, Raza MA, Wu X, Wang Z, Liu Q, Wang R, Wang X, Yong T (2018) Effect of shading and light recovery on the growth, leaf structure, and photosynthetic performance of soybean in a maize-soybean relay-strip intercropping system. PLoS ONE 13(5)
FAO (2017) The future of food and agriculture—trends and challenges. Rome
Farber A, Young AJ, Ruban AV, Horton P, Jahns P (1997) Dynamics of xanthophyll-cycle activity in different antenna subcomplexes in the photosynthetic membranes of higher plants (the relationship between zeaxanthin conversion and nonphotochemical fluorescence quenching). Plant Physiol 115(4):1609–1618. https://doi.org/10.1104/pp.115.4.1609
Feild TS, Lee DW, Holbrook NM (2001) Why leaves turn red in autumn. The role of anthocyanins in senescing leaves of red-osier dogwood. Plant Physiol 127(2):566–574
Feng L, Raza MA, Li Z, Chen Y, Khalid MHB, Du J, Liu W, Wu X, Song C, Yu L, Zhang Z, Yuan S, Yang W, Yang F (2019) The influence of light intensity and leaf movement on photosynthesis characteristics and carbon balance of soybean. Front Plant Sci. https://doi.org/10.3389/fpls.2018.01952
Article PubMed PubMed Central Google Scholar
Frank HA, Cogdell RJ (1996) Carotenoids in photosynthesis. Photochem Photobiol 63(3):257–264. https://doi.org/10.1111/j.1751-1097.1996.tb03022.x
Friedland N, Negi S, Vinogradova-Shah T, Wu G, Ma L, Flynn S, Kumssa T, Lee C-H, Sayre R (2019) Fine-tuning the photosynthetic light harvesting apparatus for improved photosynthetic efficiency and biomass yield. Sci Rep 9(1):1–12
Fryer MJ, Oxborough K, Mullineaux PM, Baker NR (2002) Imaging of photo-oxidative stress responses in leaves. J Exp Bot 53(372):1249–1254
CAS PubMed Google Scholar
Gale MD, Youssefian S (1985) Dwarfing genes in wheat. Progress in plant breeding. Butterworth and Co, London
Gan F, Zhang S, Rockwell NC, Martin SS, Lagarias JC, Bryant DA (2014) Extensive remodeling of a cyanobacterial photosynthetic apparatus in far-red light. Science 345(6202):1312–1317. https://doi.org/10.1126/science.1256963
Garcia-Molina A, Leister D (2020) Accelerated relaxation of photoprotection impairs biomass accumulation in Arabidopsis. Nature Plants 6(1):9–12
Article PubMed Google Scholar
García-Plazaola JI, Hormaetxe K, Hernández A, Olano JM, Becerril JM (2004) The lutein epoxide cycle in vegetative buds of woody plants. Funct Plant Biol 31(8):815–823. https://doi.org/10.1071/FP04054
García-Plazaola JI, Matsubara S, Osmond CB (2007) The lutein epoxide cycle in higher plants: its relationships to other xanthophyll cycles and possible functions. Funct Plant Biol 34(9):759–773
Garrido JL, Zapata M, Muñiz S (1995) Spectral characterization of new chlorophyll c pigments isolated from emiliania huxleyi (prymnesiophyceae) by high-performance liquid chromatography. J Phycol 31(5):761–768. https://doi.org/10.1111/j.0022-3646.1995.00761.x
Giberti S, Giovannini D, Forlani G (2019) Carotenoid cleavage in chromoplasts of white and yellow-fleshed peach varieties. J Sci Food Agric 99(4):1795–1803
Girolomoni L, Cazzaniga S, Pinnola A, Perozeni F, Ballottari M, Bassi R (2019) LHCSR3 is a nonphotochemical quencher of both photosystems in Chlamydomonas reinhardtii . Proc Natl Acad Sci 116(10):4212–4217
Giroux R, Berinstain A, Braham S, Graham T, Bamsey M, Boyd K, Silver M, Lussier-Desbiens A, Lee P, Boucher M (2006) Greenhouses in extreme environments: the Arthur Clarke Mars Greenhouse design and operation overview. Adv Space Res 38(6):1248–1259
Glowacka K, Kromdijk J, Kucera K, Xie J, Cavanagh AP, Leonelli L, Leakey ADB, Ort DR, Niyogi KK, Long SP (2018) Photosystem II subunit S overexpression increases the efficiency of water use in a field-grown crop. Nat Commun 9(1):868. https://doi.org/10.1038/s41467-018-03231-x
Goss R (2003) Substrate specificity of the violaxanthin de-epoxidase of the primitive green alga Mantoniella squamata (Prasinophyceae). Planta 217(5):801–812
Gotoh E, Suetsugu N, Yamori W, Ishishita K, Kiyabu R, Fukuda M, Higa T, Shirouchi B, Wada M (2018) Chloroplast accumulation response enhances leaf photosynthesis and plant biomass production. Plant Physiol 178(3):1358–1369
Gould KS, Jay-Allemand C, Logan BA, Baissac Y, Bidel LP (2018) When are foliar anthocyanins useful to plants? Re-evaluation of the photoprotection hypothesis using Arabidopsis thaliana mutants that differ in anthocyanin accumulation. Environ Exp Bot 154:11–22
Granick S (1965) Evolution of heme and chlorophyll. In: Bryson G, Vogel H (eds) Evolving genes and proteins. Academic Press, New York, pp 67–88
Gruszecki WI, Strzałka K (2005) Carotenoids as modulators of lipid membrane physical properties. Biochim Biophys Acta 1740(2):108–115. https://doi.org/10.1016/j.bbadis.2004.11.015
Gu J, Zhou Z, Li Z, Chen Y, Wang Z, Zhang H (2017) Rice ( Oryza sativa L.) with reduced chlorophyll content exhibit higher photosynthetic rate and efficiency, improved canopy light distribution, and greater yields than normally pigmented plants. Field Crops Res. https://doi.org/10.1016/j.fcr.2016.10.008
Hashimoto H, Uragami C, Cogdell RJ (2016) Carotenoids and photosynthesis. Carotenoids in nature. Springer, Berlin, pp 111–139
Haupt W, Scheuerlein R (1990) Chloroplast movement. Plant Cell Environ 13(7):595–614. https://doi.org/10.1111/j.1365-3040.1990.tb01078.x
Havaux M (1998) Carotenoids as membrane stabilisers in chloroplasts. Trends Plant Sci 3:147–151
Havaux M, Dall’Osto L, Bassi R (2007) Zeaxanthin has enhanced antioxidant capacity with respect to all other xanthophylls in Arabidopsis leaves and functions independent of binding to PSII antennae. Plant Physiol 145(4):1506–1520
Hieber AD, Bugos RC, Yamamoto HY (2000) Plant lipocalins: violaxanthin de-epoxidase and zeaxanthin epoxidase. Biochim Biophys Acta 1482(1–2):84–91
Holt TK, Krogmann DW (1981) A carotenoid protein from cyanobacteria. Biochim Biophys Acta 637:408–414
Hughes NM, Lev-Yadun S (2015) Red/purple leaf margin coloration: potential ecological and physiological functions. Environ Exp Bot 119:27–39
Hughes NM, Smith WK (2007) Seasonal photosynthesis and anthocyanin production in 10 broadleaf evergreen species. Funct Plant Biol 34(12):1072–1079
Jahns P, Wehner A, Paulsen H, Hobe S (2001) De-epoxidation of violaxanthin after reconstitution into different carotenoid binding sites of light-harvesting complex II. J Biol Chem 276(25):22154–22159
Jahns P, Latowski D, Strzalka K (2009) Mechanism and regulation of the violaxanthin cycle: the role of antenna proteins and membrane lipids. Biochim Biophys Acta 1787(1):3–14
Jeffrey S (1989) Chlorophyll pigments and their distribution in the chromophyte algae. Chrompophyte Algae 13–36
Jia H, Förster B, Chow WS, Pogson BJ, Osmond CB (2013) Decreased photochemical efficiency of photosystem II following sunlight exposure of shade-grown leaves of avocado: because of, or in spite of, two kinetically distinct xanthophyll cycles? Plant Physiol 161(2):836–852
Jiang Y, Hu H, Ma Y, Zhou J (2020) Genome-wide identification and characterization of the fibrillin gene family in Triticum aestivum. PeerJ 8:e9225
Jin H, Li M, Duan S, Fu M, Dong X, Liu B, Feng D, Wang J, Wang H-B (2016) Optimization of light-harvesting pigment improves photosynthetic efficiency. Plant Physiol 172(3):1720–1731
Johnson MP, Ruban AV (2010) Arabidopsis plants lacking PsbS protein possess photoprotective energy dissipation. Plant J 61(2):283–289
Johnson MP, Havaux M, Triantaphylides C, Ksas B, Pascal AA, Robert B, Davison PA, Ruban AV, Horton P (2007) Elevated zeaxanthin bound to oligomeric LHCII enhances the resistance of Arabidopsis to photooxidative stress by a lipid-protective, antioxidant mechanism. J Biol Chem 282(31):22605–22618. https://doi.org/10.1074/jbc.M702831200
Johnson MP, Davison PA, Ruban AV, Horton P (2008) The xanthophyll cycle pool size controls the kinetics of non-photochemical quenching in Arabidopsis thaliana . FEBS Lett 582(2):262–266
Jordheim M, Calcott K, Gould KS, Davies KM, Schwinn KE, Andersen ØM (2016) High concentrations of aromatic acylated anthocyanins found in cauline hairs in Plectranthus ciliatus . Phytochemistry 128:27–34
Josse EM, Simkin AJ, Gaffe J, Laboure AM, Kuntz M, Carol P (2000) A plastid terminal oxidase associated with carotenoid desaturation during chromoplast differentiation. Plant Physiol 123(4):1427–1436
Kalve S, Fotschki J, Beeckman T, Vissenberg K, Beemster GT (2014) Three-dimensional patterns of cell division and expansion throughout the development of Arabidopsis thaliana leaves. J Exp Bot 65(22):6385–6397
Kanazawa A, Kramer DM (2002) In vivo modulation of nonphotochemical exciton quenching (NPQ) by regulation of the chloroplast ATP synthase. Proc Natl Acad Sci 99(20):12789–12794. https://doi.org/10.1073/pnas.182427499
Kasahara M, Kagawa T, Oikawa K, Suetsugu N, Miyao M, Wada M (2002) Chloroplast avoidance movement reduces photodamage in plants. Nature 420(6917):829
Kashiyama Y, Miyashita H, Ohkubo S, Ogawa NO, Chikaraishi Y, Takano Y, Suga H, Toyofuku T, Nomaki H, Kitazato H, Nagata T, Ohkouchi N (2008) Evidence of global chlorophyll d . Science 321(5889):658. https://doi.org/10.1126/science.1158761
Kerfeld CA (2004a) Structure and function of the water-soluble carotenoid-binding proteins of cyanobacteria. Photosynth Res 81(3):215–225. https://doi.org/10.1023/B:PRES.0000036886.60187.c8
Kerfeld CA (2004b) Water-soluble carotenoid proteins of cyanobacteria. Arch Biochem Biophys 430(1):2–9. https://doi.org/10.1016/j.abb.2004.03.018
Kim E-H, Lee Y, Kim HU (2015) Fibrillin 5 is essential for plastoquinone-9 biosynthesis by binding to solanesyl diphosphate synthases in Arabidopsis. Plant Cell 27(10):2956–2971. https://doi.org/10.1105/tpc.15.00707
Kirilovsky D (2007) Photoprotection in cyanobacteria: the orange carotenoid protein (OCP)-related non-photochemical-quenching mechanism. Photosynth Res 93(1–3):7–16. https://doi.org/10.1007/s11120-007-9168-y
Kirilovsky D, Kerfeld CA (2012) The orange carotenoid protein in photoprotection of photosystem II in cyanobacteria. Biochim Biophys Acta 1:158–166
Kirk JTO (2011) Light and photosynthesis in aquatic ecosystems. Cambridge University Press, Cambridge
Koch TC, Goldman IL (2005) Relationship of carotenoids and tocopherols in a sample of carrot root-color accessions and carrot germplasm carrying Rp and rp alleles. J Agric Food Chem 53(2):325–331. https://doi.org/10.1021/jf048272z
Kodama Y, Tsuboi H, Kagawa T, Wada M (2008) Low temperature-induced chloroplast relocation mediated by a blue light receptor, phototropin 2, in fern gametophytes. J Plant Res 121(4):441–448. https://doi.org/10.1007/s10265-008-0165-9
Komatsu A, Terai M, Ishizaki K, Suetsugu N, Tsuboi H, Nishihama R, Yamato KT, Wada M, Kohchi T (2014) Phototropin encoded by a single-copy gene mediates chloroplast photorelocation movements in the liverwort Marchantia polymorpha . Plant Physiol 166(1):411–427. https://doi.org/10.1104/pp.114.245100
Krieger-Liszkay A, Fufezan C, Trebst A (2008) Singlet oxygen production in photosystem II and related protection mechanism. Photosynth Res 98(1–3):551–564. https://doi.org/10.1007/s11120-008-9349-3
Kromdijk J, Glowacka K, Leonelli L, Gabilly ST, Iwai M, Niyogi KK, Long SP (2016) Improving photosynthesis and crop productivity by accelerating recovery from photoprotection. Science 354(6314):857–861. https://doi.org/10.1126/science.aai8878
Kume A, Akitsu T, Nasahara KN (2018) Why is chlorophyll b only used in light-harvesting systems? J Plant Res 131(6):961–972. https://doi.org/10.1007/s10265-018-1052-7
Langenkämper G, Manac’h N, Broin M, Cuiné S, Becuwe N, Kuntz M, Rey P (2001) Accumulation of plastid lipid-associated proteins (fibrillin/CDSP34) upon oxidative stress, ageing and biotic stress in Solanaceae and in response to drought in other species. J Exp Bot 52(360):1545–1554. https://doi.org/10.1093/jexbot/52.360.1545
Latowski D, Burda K, Strzalka K (2000) A mathematical model describing kinetics of conversion of violaxanthin to zeaxanthin via intermediate antheraxanthin by the xanthophyll cycle enzyme violaxanthin de-epoxidase. J Theor Biol 206(4):507–514. https://doi.org/10.1006/jtbi.2000.2141
Latowski D, Kuczynska P, Strzalka K (2011) Xanthophyll cycle—a mechanism protecting plants against oxidative stress. Redox Rep 16(2):78–90. https://doi.org/10.1179/174329211X13020951739938
Ledford HK, Niyogi KK (2005) Singlet oxygen and photo-oxidative stress management in plants and algae. Plant Cell Environ 28:1037–1045
Lefebvre S, Lawson T, Fryer M, Zakhleniuk OV, Lloyd JC, Raines CA (2005) Increased sedoheptulose-1,7-bisphosphatase activity in transgenic tobacco plants stimulates photosynthesis and growth from an early stage in development. Plant Physiol 138(1):451–460. https://doi.org/10.1104/pp.104.055046
Li Z, Ahn TK, Avenson TJ, Ballottari M, Cruz JA, Kramer DM, Bassi R, Fleming GR, Keasling JD, Niyogi KK (2009) Lutein accumulation in the absence of zeaxanthin restores nonphotochemical quenching in the Arabidopsis thaliana npq1 mutant. Plant Cell 21(6):1798–1812
Lin J, Rouseff RL, Barros S, Naim M (2002) Aroma composition changes in early season grapefruit juice produced from thermal concentration. J Agric Food Chem 50(4):813–819
Liu H, Zhang H, Niedzwiedzki DM, Prado M, He G, Gross ML, Blankenship RE (2013) Phycobilisomes supply excitations to both photosystems in a megacomplex in cyanobacteria. Science 342(6162):1104–1107
López-Calcagno PE, Fisk SJ, Brown K, Bull SE, P.F. S, Raines CA, (2018) Overexpressing the H-protein of the glycine cleavage system increases biomass yield in glasshouse and field-grown transgenic tobacco plants. Plant Biotechnol J 17(1):141–151. https://doi.org/10.1111/pbi.12953
Lopez-Calcagno PE, Brown KL, Simkin AJ, Fisk SJ, Lawson T, Raines CA (2020) Simultaneous stimulation of RuBP regeneration and electron transport increases productivity and water use efficiency under field conditions. bioRxiv
Lopez-Obando M, Ligerot Y, Bonhomme S, Boyer F-D, Rameau C (2015) Strigolactone biosynthesis and signaling in plant development. Development 142(21):3615–3619. https://doi.org/10.1242/dev.120006
Lovelock CE, Clough BF, Woodrow IE (1992) Distribution and accumulation of ultraviolet-radiation-absorbing compounds in leaves of tropical mangroves. Planta 188(2):143–154
Mahattanatawee K, Rouseff R, Valim MF, Naim M (2005) Identification and aroma impact of norisoprenoids in orange juice. J Agric Food Chem 53(2):393–397
Matsubara S, Naumann M, Martin R, Nichol C, Rascher U, Morosinotto T, Bassi R, Osmond B (2005) Slowly reversible de-epoxidation of lutein-epoxide in deep shade leaves of a tropical tree legume may ‘lock-in’ lutein-based photoprotection during acclimation to strong light. J Exp Bot 56(411):461–468. https://doi.org/10.1093/jxb/eri012
Matsubara S, Morosinotto T, Osmond CB, Bassi R (2007) Short-and long-term operation of the lutein-epoxide cycle in light-harvesting antenna complexes. Plant Physiol 144(2):926–941
Matsubara S, Krause GH, Aranda J, Virgo A, Beisel KG, Jahns P, Winter K (2009) Sun-shade patterns of leaf carotenoid composition in 86 species of neotropical forest plants. Funct Plant Biol 36(1):20–36
Matsubara S, Förster B, Waterman M, Robinson SA, Pogson BJ, Gunning B, Osmond B (2012) From ecophysiology to phenomics: some implications of photoprotection and shade–sun acclimation in situ for dynamics of thylakoids in vitro. Philos Trans R Soc B 367(1608):3503–3514
Merzlyak MN, Chivkunova OB, Solovchenko AE, Naqvi KR (2008) Light absorption by anthocyanins in juvenile, stressed, and senescing leaves. J Exp Bot 59(14):3903–3911
Mielke SP, Kiang NY, Blankenship RE, Gunner MR, Mauzerall D (2011) Efficiency of photosynthesis in a Chl d -utilizing cyanobacterium is comparable to or higher than that in Chl a -utilizing oxygenic species. Biochim Biophys Acta 9:1231–1236. https://doi.org/10.1016/j.bbabio.2011.06.007
Millenaar FF, Van Zanten M, Cox MC, Pierik R, Voesenek LA, Peeters AJ (2009) Differential petiole growth in Arabidopsis thaliana : photocontrol and hormonal regulation. New Phytol 184(1):141–152
Mimuro M, Akimoto S, Gotoh T, Yokono M, Akiyama M, Tsuchiya T, Miyashita H, Kobayashi M, Yamazaki I (2004) Identification of the primary electron donor in PS II of the Chl d-dominated cyanobacterium Acaryochloris marina . FEBS Lett 556(1–3):95–98. https://doi.org/10.1016/s0014-5793(03)01383-8
Mirkovic T, Ostroumov EE, Anna JM, van Grondelle R, Govindjee SGD (2017) Light absorption and energy transfer in the antenna complexes of photosynthetic organisms. Chem Rev 117(2):249–293. https://doi.org/10.1021/acs.chemrev.6b00002
Miyashita H, Ikemoto H, Kurano N, Adachi K, Chihara M, Miyachi S (1996) Chlorophyll d as a major pigment. Nature 383(6599):402
Miyashita H, Ohkubo S, Komatsu H, Sorimachi Y, Fukayama D, Fujinuma D, Akutsu S, Kobayashi M (2014) Discovery of Chlorophyll d in Acaryochloris marina and Chlorophyll f in a unicellular Cyanobacterium, strain KC1, isolated from Lake Biwa. J Phys Chem Biophys 4(4):149
Mukherjee S, Stasolla C, Brûlé-Babel A, Ayele BT (2015) Isolation and characterization of rubisco small subunit gene promoter from common wheat ( Triticum aestivum L.). Plant Signal Behav 10(2):e989033
Müller P, Li X-P, Niyogi KK (2001) Non-photochemical quenching. A response to excess light energy. Plant Physiol 125(4):1558–1566
Mullineaux CW, Emlyn-Jones D (2005) State transitions: an example of acclimation to low-light stress. J Exp Bot 56(411):389–393
Niyogi KK, Li X-P, Rosenberg V, Jung H-S (2004) Is PsbS the site of non-photochemical quenching in photosynthesis? J Exp Bot 56(411):375–382. https://doi.org/10.1093/jxb/eri056
Nobel PS (2009) Chapter 5—photochemistry of photosynthesis. In: Nobel PS (ed) Physicochemical and environmental plant physiology, 4th edn. Academic Press, San Diego, pp 228–275
Nürnberg DJ, Morton J, Santabarbara S, Telfer A, Joliot P, Antonaru LA, Ruban AV, Cardona T, Krausz E, Boussac A (2018) Photochemistry beyond the red limit in chlorophyll f–containing photosystems. Science 360(6394):1210–1213
Article PubMed CAS Google Scholar
Ort DR, Melis A (2011) Optimizing antenna size to maximize photosynthetic efficiency. Plant Physiol 155(1):79–85. https://doi.org/10.1104/pp.110.165886
Ort DR, Merchant SS, Alric J, Barkan A, Blankenship RE, Bock R, Croce R, Hanson MR, Hibberd JM, Long SP (2015a) Redesigning photosynthesis to sustainably meet global food and bioenergy demand. Proc Natl Acad Sci 112(28):8529–8536
Ort DR, Merchant SS, Alric J, Barkan A, Blankenship RE, Bock R, Croce R, Hanson MR, Hibberd JM, Long SP, Moore TA, Moroney J, Niyogi KK, Parry MAJ, Peralta-Yahya PP, Prince RC, Redding KE, Spalding MH, van Wijk KJ, Vermaas WFJ, von Caemmerer S, Weber APM, Yeates TO, Yuan JS, Zhu XG (2015b) Redesigning photosynthesis to sustainably meet global food and bioenergy demand. Proc Natl Acad Sci 112(28):8529–8536. https://doi.org/10.1073/pnas.1424031112
Parry AD, Babiano MJ, Horgan R (1990) The role of cis-carotenoids in abscisic acid biosynthesis. Planta 182:118–128
Parry MA, Reynolds M, Salvucci ME, Raines C, Andralojc PJ, Zhu XG, Price GD, Condon AG, Furbank RT (2011) Raising yield potential of wheat. II. Increasing photosynthetic capacity and efficiency. J Exp Bot 62(2):453–467. https://doi.org/10.1093/jxb/erq304
Parson WW, Green BR, Van Amerongen H, Dekker J (2003) Light-harvesting antennas in photosynthesis. Kluwer Academic
Pastenes C, Pimentel P, Lillo J (2005) Leaf movements and photoinhibition in relation to water stress in field-grown beans. J Exp Bot 56(411):425–433. https://doi.org/10.1093/jxb/eri061
Perrine Z, Negi S, Sayre RT (2012) Optimization of photosynthetic light energy utilization by microalgae. Algal Res 1(2):134–142
Pettigrew WT, Hesketh JD, Peters DB, Woolley JT (1989) Characterization of canopy photosynthesis of chlorophyll-deficient soybean isolines. Crop Sci. https://doi.org/10.2135/cropsci1989.0011183X002900040040x
Pharis RP, King RW (1985) Gibberellins and reproductive development in seed plants. Annu Rev Plant Physiol 36(1):517–568
Piccolo EL, Landi M, Massai R, Remorini D, Guidi L (2020) Girled-induced anthocyanin accumulation in red-leafed Prunus cerasifera: effect on photosynthesis, photoprotection and sugar metabolism. Plant Sci 294:110456
Pieters AJ, El Souki S (2005) Effects of drought during grain filling on PS II activity in rice. J Plant Physiol 162(8):903–911. https://doi.org/10.1016/j.jplph.2004.11.001
Pieters AJ, Tezara W, Herrera A (2003) Operation of the xanthophyll cycle and degradation of D1 protein in the inducible CAM plant, Talinum triangulare , under water deficit. Ann Bot 92(3):393–399. https://doi.org/10.1093/aob/mcg153
Pietrini F, Iannelli M, Massacci A (2002) Anthocyanin accumulation in the illuminated surface of maize leaves enhances protection from photo-inhibitory risks at low temperature, without further limitation to photosynthesis. Plant Cell Environ 25(10):1251–1259
Poisson L, Schieberle P (2008) Characterization of the most odor-active compounds in an American Bourbon whisky by application of the aroma extract dilution analysis. J Agric Food Chem 56(14):5813–5819
Poulet L, Massa G, Morrow R, Bourget C, Wheeler R, Mitchell C (2014) Significant reduction in energy for plant-growth lighting in space using targeted LED lighting and spectral manipulation. Life Sci Space Res 2:43–53
Qin X, Zeevaart JAD (1999) The 9-cis-epoxycarotenoid cleavage reaction is the key regulatory step of abscisic acid biosynthesis in water-stressed bean. Proc Natl Acad Sci 96(26):15354–15361. https://doi.org/10.1073/pnas.96.26.15354
Rabinowitch HD, Budowski P, Kedar N (1975) Carotenoids and epoxide cycles in mature-green tomatoes. Planta 122(1):91–97. https://doi.org/10.1007/BF00385408
Raines CA (2003) The Calvin cycle revisited. Photosynth Res 75(1):1–10. https://doi.org/10.1023/A:1022421515027
Raines CA, Cavanagh AP, Simkin AJ (2022) Chapter 9. Improving carbon fixation. In: Ruban A, Murchie E, Foyer C (eds) Photosynthesis in action, 1st edn. Academic Press, New York
Ramel F, Birtic S, Ginies C, Soubigou-Taconnat L, Triantaphylidès C, Havaux M (2012) Carotenoid oxidation products are stress signals that mediate gene responses to singlet oxygen in plants. Proc Natl Acad Sci 109(14):5535–5540. https://doi.org/10.1073/pnas.1115982109
Ray DK, Ramankutty N, Mueller ND, West PC, Foley JA (2012) Recent patterns of crop yield growth and stagnation. Nat Commun 3(1):1–7
Ray DK, Mueller ND, West PC, Foley JA (2013) Yield trends are insufficient to double global crop production by 2050. PLoS ONE 8(6):e66428
Raymond J, Blankenship RE (2004) Biosynthetic pathways, gene replacement and the antiquity of life. Geobiology 2:199–203
Renger T, Schlodder E (2008) The primary electron donor of Photosystem II of the Cyanobacterium Acaryochloris marina is a Chlorophyll d and the water oxidation is driven by a Chlorophyll a /Chlorophyll d heterodimer. J Phys Chem B 112(25):7351–7354. https://doi.org/10.1021/jp801900e
Rey P, Gillet B, Romer S, Eymery F, Massimino J, Peltier G, Kuntz M (2000) Over-expression of a pepper plastid lipid-associated protein in tobacco leads to changes in plastid ultrastructure and plant development upon stress. Plant J 21(5):483–494
Reynolds M, Foulkes MJ, Slafer GA, Berry P, Parry MAJ, Snape JW, Angus WJ (2009) Raising yield potential in wheat. J Exp Bot 60(7):1899–1918. https://doi.org/10.1093/jxb/erp016
RSOL (2009) Royal Society of London. Science and the Sustainable Intensification of Global Agriculture (Royal Society, London, Reaping the Benefits, p 2009
Ruban AV, Young AJ, Pascal AA, Horton P (1994) The effects of illumination on the xanthophyll composition of the Photosystem II light-harvesting complexes of spinach thylakoid membranes. Plant Physiol 104(1):227–234. https://doi.org/10.1104/pp.104.1.227
Rubio A, Rambla JL, Santaella M, Gomez MD, Orzaez D, Granell A, Gomez-Gomez L (2008) Cytosolic and plastoglobule-targeted carotenoid dioxygenases from Crocus sativus are both involved in beta-ionone release. J Biol Chem 283(36):24816–24825. https://doi.org/10.1074/jbc.M804000200
Sacharz J, Giovagnetti V, Ungerer P, Mastroianni G, Ruban AV (2017) The xanthophyll cycle affects reversible interactions between PsbS and light-harvesting complex II to control non-photochemical quenching. Nat Plants 3(2):1–9
Sakowska K, Alberti G, Genesio L, Peressotti A, Delle Vedove G, Gianelle D, Colombo R, Rodeghiero M, Panigada C, Juszczak R (2018) Leaf and canopy photosynthesis of a chlorophyll deficient soybean mutant. Plant Cell Environ 41(6):1427–1437
Sandmann G (1989) Inhibition of carotenoid biosynthesis by herbicides. Target sites of herbicides action
Sankari M, Rao PR, Hemachandran H, Pullela PK, Tayubi IA, Subramanian B, Gothandam K, Singh P, Ramamoorthy S (2018) Prospects and progress in the production of valuable carotenoids: insights from metabolic engineering, synthetic biology, and computational approaches. J Biotechnol 266:89–101
Scheer H (1991) Chemistry of chlorophylls. Chlorophylls 3–30
Schwartz SH, Qin X, Zeevaart JA (2001) Characterization of a novel carotenoid cleavage dioxygenase from plants. J Biol Chem 276(27):25208–25211. https://doi.org/10.1074/jbc.M102146200
Schwartz SH, Qin X, Loewen MC (2004) The biochemical characterization of two carotenoid cleavage enzymes from Arabidopsis indicates that a carotenoid-derived compound inhibits lateral branching. J Biol Chem 279(45):46940–46945
Senge M, Ryan A, Letchford K, MacGowan S, Mielke T (2014) Chlorophylls, symmetry, chirality, and photosynthesis. Symmetry 6(3):781–843
Siefermann-Harms D (1994) Light and temperature control of season-dependent changes in the α- and β-carotene content of spruce needles. J Plant Physiol 143(4):488–494. https://doi.org/10.1016/S0176-1617(11)81811-0
Simkin AJ (2019) Genetic engineering for global food security: photosynthesis and biofortification. Plants 8(12):586
Article CAS PubMed Central Google Scholar
Simkin AJ (2021) Carotenoids and apocarotenoids in planta: their role in plant development, contribution to the flavour and aroma of fruits and flowers, and their nutraceutical benefits. Plants 10(11):2321
Simkin AJ, Breitenbach J, Kuntz M, Sandmann G (2000) In vitro and in situ inhibition of carotenoid biosynthesis in Capsicum annuum by bleaching herbicides. J Agric Food Chem 48(10):4676–4680
Simkin AJ, Laizet Y, Kuntz M (2004a) Plastid lipid associated proteins of the fibrillin family: structure, localisation, functions and gene expression. In: Pandalai SG (ed) Recent research developmetns in biochemistry, vol 5. Research Signpost, India, pp 307–316
Simkin AJ, Schwartz SH, Auldridge M, Taylor MG, Klee HJ (2004b) The tomato carotenoid cleavage dioxygenase 1 genes contribute to the formation of the flavor volatiles β-ionone, pseudoionone, and geranylacetone. Plant J 40(6):882–892
Simkin AJ, Underwood BA, Auldridge M, Loucas HM, Shibuya K, Schmelz E, Clark DG, Klee HJ (2004c) Circadian regulation of the PhCCD1 carotenoid cleavage dioxygenase controls emission of β-ionone, a fragrance volatile of petunia flowers. Plant Physiol 136(3):3504–3514
Simkin AJ, Gaffé J, Alcaraz J-P, Carde J-P, Bramley PM, Fraser PD, Kuntz M (2007a) Fibrillin influence on plastid ultrastructure and pigment content in tomato fruit. Phytochemistry 68(11):1545–1556
Simkin AJ, Gaffe J, Alcaraz JP, Carde JP, Bramley PM, Fraser PD, Kuntz M (2007b) Fibrillin influence on plastid ultrastructure and pigment content in tomato fruit. Phytochemistry 68(11):1545–1556. https://doi.org/10.1016/j.phytochem.2007.03.014
Simkin AJ, Moreau H, Kuntz M, Pagny G, Lin C, Tanksley S, McCarthy J (2008) An investigation of carotenoid biosynthesis in Coffea canephora and Coffea arabica . J Plant Physiol 165(10):1087–1106. https://doi.org/10.1016/j.jplph.2007.06.016
Simkin AJ, Kuntz M, Moreau H, McCarthy J (2010) Carotenoid profiling and the expression of carotenoid biosynthetic genes in developing coffee grain. Plant Physiol Biochem 48(6):434–442. https://doi.org/10.1016/j.plaphy.2010.02.007
Simkin AJ, McAusland L, Headland LR, Lawson T, Raines CA (2015) Multigene manipulation of photosynthetic carbon assimilation increases CO 2 fixation and biomass yield in tobacco. J Exp Bot 66(13):4075–4090. https://doi.org/10.1093/jxb/erv204
Simkin AJ, Lopez-Calcagno PE, Davey PA, Headland LR, Lawson T, Timm S, Bauwe H, Raines CA (2017a) Simultaneous stimulation of sedoheptulose 1, 7-bisphosphatase, fructose 1, 6-bisphophate aldolase and the photorespiratory glycine decarboxylase-H protein increases CO 2 assimilation, vegetative biomass and seed yield in Arabidopsis. Plant Biotechnol J 15(7):805–816
Simkin AJ, McAusland L, Lawson T, Raines CA (2017b) Over-expression of the RieskeFeS protein increases electron transport rates and biomass yield. Plant Physiol 175:134–145. https://doi.org/10.1101/133702
Simkin AJ, Lopez-Calcagno PE, Raines CA (2019) Feeding the world: improving photosynthetic efficiency for sustainable crop production. J Exp Bot 70(4):1119–1140. https://doi.org/10.1093/jxb/ery445
Simkin AJ, Faralli M, Ramamoorthy S, Lawson T (2020) Photosynthesis in non-foliar tissues: implications for yield. Plant J 101(4):1001–1015
Singh DK, McNellis TW (2011) Fibrillin protein function: the tip of the iceberg? Trends Plant Sci 16(8):432–441. https://doi.org/10.1016/j.tplants.2011.03.014
Singh DK, Maximova SN, Jensen PJ, Lehman BL, Ngugi HK, McNellis TW (2010) FIBRILLIN4 is required for plastoglobule development and stress resistance in apple and Arabidopsis. Plant Physiol 154(3):1281–1293. https://doi.org/10.1104/pp.110.164095
Siva R (2007) Status of natural dyes and dye-yielding plants in India. Curr Sci 92(7):00113891
Slattery RA, VanLoocke A, Bernacchi CJ, Zhu X-G, Ort DR (2017) Photosynthesis, light use efficiency, and yield of reduced-chlorophyll soybean mutants in field conditions. Front Plant Sci 8:549
Snowden KC, Simkin AJ, Janssen BJ, Templeton KR, Loucas HM, Simons JL, Karunairetnam S, Gleave AP, Clark DG, Klee HJ (2005) The decreased apical dominance1/Petunia hybrida CAROTENOID CLEAVAGE DIOXYGENASE8 gene affects branch production and plays a role in leaf senescence, root growth, and flower development. Plant Cell 17(3):746–759
Sobiechowska-Sasim M, Ston-Egiert J, Kosakowska A (2014) Quantitative analysis of extracted phycobilin pigments in cyanobacteria-an assessment of spectrophotometric and spectrofluorometric methods. J Appl Phycol 26(5):2065–2074. https://doi.org/10.1007/s10811-014-0244-3
Solovchenko A, Merzlyak M (2008) Screening of visible and UV radiation as a photoprotective mechanism in plants. Russ J Plant Physiol 55(6):719
Songaila E, Augulis RN, Gelzinis A, Butkus V, Gall A, Büchel C, Robert B, Zigmantas D, Abramavicius D, Valkunas L (2013) Ultrafast energy transfer from chlorophyll c2 to chlorophyll a in fucoxanthin–chlorophyll protein complex. J Phys Chem Lett 4:3590–3595
Standfuss J, Terwisscha van Scheltinga AC, Lamborghini M, Kuhlbrandt W (2005) Mechanisms of photoprotection and nonphotochemical quenching in pea light-harvesting complex at 2.5 A resolution. EMBO J 24 (5):919–928. doi: https://doi.org/10.1038/sj.emboj.7600585
Steyn WJ, Wand S, Holcroft D, Jacobs G (2002) Anthocyanins in vegetative tissues: a proposed unified function in photoprotection. New Phytol 155(3):349–361
Takahashi S, Badger MR (2011) Photoprotection in plants: a new light on photosystem II damage. Trends Plant Sci 16(1):53–60
Takizawa K, Kanazawa A, Kramer DM (2008) Depletion of stromal P i induces high “energy-dependent” antenna exciton quenching q E by decreasing proton conductivity at CF O -CF 1 ATP synthase. Plant Cell Environ 31(2):235–243. https://doi.org/10.1111/j.1365-3040.2007.01753.x
Tan B-C, Joseph LM, Deng W-T, Liu L, Li Q-B, Cline K, McCarty DR (2003) Molecular characterization of the Arabidopsis 9-cis epoxycarotenoid dioxygenase gene family. Plant J 35(1):44–56. https://doi.org/10.1046/j.1365-313X.2003.01786.x
Tanaka R, Kobayashi K, Masuda T (2011) Tetrapyrrole metabolism in Arabidopsis thaliana . Arabidopsis Book 9:e0145. https://doi.org/10.1199/tab.0145
Tattini M, Landi M, Brunetti C, Giordano C, Remorini D, Gould KS, Guidi L (2014) Epidermal coumaroyl anthocyanins protect sweet basil against excess light stress: multiple consequences of light attenuation. Physiol Plant 152(3):585–598
Thayer SS, Björkman O (1990) Leaf xanthophyll content and composition in sun and shade determined by HPLC. Photosynth Res 23(3):331–343
Tilman D, Clark M (2015) Food, agriculture and the environment: can we feed the world and save the Earth? Daedalus 144:8–23
Timm S, Florian A, Arrivault S, Stitt M, Fernie AR, Bauwe H (2012) Glycine decarboxylase controls photosynthesis and plant growth. FEBS Lett 586(20):3692–3697. https://doi.org/10.1016/j.febslet.2012.08.027
Tlałka M, Runquist M, Fricker M (1999) Light perception and the role of the xanthophyll cycle in blue-light-dependent chloroplast movements in Lemna trisulca L. Plant J 20(4):447–459
Verhoeven A (2014) Sustained energy dissipation in winter evergreens. New Phytol 201(1):57–65
Vialet-Chabrand S, Matthews JS, Simkin AJ, Raines CA, Lawson T (2017) Importance of fluctuations in light on plant photosynthetic acclimation. Plant Physiol 173(4):2163–2179
Vishnevetsky M, Ovadis M, Vainstein A (1999) Carotenoid sequestration in plants: the role of carotenoid-associated proteins. Trends Plant Sci 4(6):232–235
Vogel JT, Walter MH, Giavalisco P, Lytovchenko A, Kohlen W, Charnikhova T, Simkin AJ, Goulet C, Strack D, Bouwmeester HJ (2010) SlCCD7 controls strigolactone biosynthesis, shoot branching and mycorrhiza-induced apocarotenoid formation in tomato. Plant J 61(2):300–311
Voitsekhovskaja O, Tyutereva E (2015) Chlorophyll b in angiosperms: functions in photosynthesis, signaling and ontogenetic regulation. J Plant Physiol 189:51–64
Wada M, Kong SG (2011) Analysis of chloroplast movement and relocation in Arabidopsis. Methods Mol Biol 774:87–102. https://doi.org/10.1007/978-1-61779-234-2_6
Walters RG (2005) Towards an understanding of photosynthetic acclimation. J Exp Bot 56(411):435–447. https://doi.org/10.1093/jxb/eri060
Wilson A, Ajlani G, Verbavatz JM, Vass I, Kerfeld CA, Kirilovsky D (2006) A soluble carotenoid protein involved in phycobilisome-related energy dissipation in cyanobacteria. Plant Cell 18(4):992–1007. https://doi.org/10.1105/tpc.105.040121
Winterhalter P, Gök R (2013) TDN and β-Damascenone: two important carotenoid metabolites in wine. Carotenoid cleavage products. ACS Publications, Washington, pp 125–137
Woitsch S, Romer S (2003) Expression of xanthophyll biosynthetic genes during light-dependent chloroplast differentiation. Plant Physiol 132(3):1508–1517. https://doi.org/10.1104/pp.102.019364
WorldBank (2008) World development report 2008: agriculture for development. World Bank, Washington
Wu YP, Krogmann DW (1997) The orange carotenoid protein of Synechocystis PCC 6803. Biochim Biophys Acta 1322(1):1–7. https://doi.org/10.1016/s0005-2728(97)00067-4
Yadav SK, Khatri K, Rathore MS, Jha B (2018) Introgression of UfCyt c 6 , a thylakoid lumen protein from a green seaweed Ulva fasciata Delile enhanced photosynthesis and growth in tobacco. Mol Biol Rep 45(6):1745–1758. https://doi.org/10.1007/s11033-018-4318-1
Yamamoto HY, Higashi RM (1978) Violaxanthin de-epoxidase: lipid composition and substrate specificity. Arch Biochem Biophys 190(2):514–522
Yamori W (2016) Photosynthetic response to fluctuating environments and photoprotective strategies under abiotic stress. J Plant Res 129(3):379–395
Yang Y, Sulpice R, Himmelbach A, Meinhard M, Christmann A, Grill E (2006) Fibrillin expression is regulated by abscisic acid response regulators and is involved in abscisic acid-mediated photoprotection. Proc Natl Acad Sci 103(15):6061–6066. https://doi.org/10.1073/pnas.0501720103
Youssef A, Laizet Y, Block MA, Marechal E, Alcaraz JP, Larson TR, Pontier D, Gaffe J, Kuntz M (2010) Plant lipid-associated fibrillin proteins condition jasmonate production under photosynthetic stress. Plant J 61(3):436–445. https://doi.org/10.1111/j.1365-313X.2009.04067.x
Zapata M, Garrido JL, Jeffrey SW (2006) Chlorophyll c pigments: current status. In: Grimm B, Porra RJ, Rüdiger W, Scheer H (eds) Chlorophylls and bacteriochlorophylls: biochemistry, biophysics, functions and applications. Springer, Dordrecht, pp 39–53
Zeidler C, Vrakking V, Bamsey M, Poulet L, Zabel P, Schubert D, Paille C, Mazzoleni E, Domurath N (2017) Greenhouse module for space system: a lunar greenhouse design. Open Agric 2(1):116–132
Zhen S, Bugbee B (2020) Far-red photons have equivalent efficiency to traditional photosynthetic photons: implications for redefining photosynthetically active radiation. Plant Cell Environ. https://doi.org/10.1111/pce.13730
Download references
Acknowledgements
The authors take this opportunity to thank the management of VIT for providing facilities to carry out this research work.
This research did not receive any specific grant from funding agencies in the public, commercial, or not-for-profit sectors. AJS is supported by the Growing Kent and Medway Programme, UK; Ref 107139.
Author information
Andrew J. Simkin and Leepica Kapoor have contributed equally to this work.
Authors and Affiliations
School of Biosciences, University of Kent, Canterbury, CT2 7NJ, United Kingdom
Andrew J. Simkin
School of Bio Sciences and Technology, Vellore Institute of Technology, Vellore, 632014, Tamil Nadu, India
Leepica Kapoor, C. George Priya Doss & Siva Ramamoorthy
School of Life Sciences, University of Essex, Wivenhoe Park, Colchester, CO4 3SQ, United Kingdom
Tracy Lawson
OSFC, Scrivener Drive, Pinewood, Ipswich, IP8 3SU, United Kingdom
Tanja A. Hofmann
You can also search for this author in PubMed Google Scholar
Contributions
AJS, TL and SR conceived the project and supervised the co-authors in writing the draft. AJS and LK wrote the article. TL, TAH and GPDC had an equal contribution in collecting information. AS, TL and SR supervised the co- authors in writing the draft, SR agrees to serve as the author responsible for contact and ensures communication.
Corresponding author
Correspondence to Siva Ramamoorthy .
Ethics declarations
Conflict of interest.
The authors have no conflicts of interest to declare that are relevant to the content of this article.
Ethical approval
The authors declare that this article does not contain any research with human or animal subjects.
Additional information
Publisher's note.
Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Rights and permissions
Reprints and permissions
About this article
Simkin, A.J., Kapoor, L., Doss, C.G.P. et al. The role of photosynthesis related pigments in light harvesting, photoprotection and enhancement of photosynthetic yield in planta. Photosynth Res 152 , 23–42 (2022). https://doi.org/10.1007/s11120-021-00892-6
Download citation
Received : 25 September 2021
Accepted : 13 December 2021
Published : 22 January 2022
Issue Date : April 2022
DOI : https://doi.org/10.1007/s11120-021-00892-6
Share this article
Anyone you share the following link with will be able to read this content:
Sorry, a shareable link is not currently available for this article.
Provided by the Springer Nature SharedIt content-sharing initiative
- Anthocyanins
- Carotenoids
- Chlorophyll
- Xanthophyll cycle
- Nonphotochemical quenching
- Phycobilins
- Find a journal
- Publish with us
- Track your research
ORIGINAL RESEARCH article
Photosynthetic physiology of blue, green, and red light: light intensity effects and underlying mechanisms.
- Horticultural Physiology Laboratory, Department of Horticulture, University of Georgia, Athens, GA, United States
Red and blue light are traditionally believed to have a higher quantum yield of CO 2 assimilation ( QY , moles of CO 2 assimilated per mole of photons) than green light, because green light is absorbed less efficiently. However, because of its lower absorptance, green light can penetrate deeper and excite chlorophyll deeper in leaves. We hypothesized that, at high photosynthetic photon flux density ( PPFD ), green light may achieve higher QY and net CO 2 assimilation rate ( A n ) than red or blue light, because of its more uniform absorption throughtout leaves. To test the interactive effects of PPFD and light spectrum on photosynthesis, we measured leaf A n of “Green Tower” lettuce ( Lactuca sativa ) under red, blue, and green light, and combinations of those at PPFD s from 30 to 1,300 μmol⋅m –2 ⋅s –1 . The electron transport rates ( J ) and the maximum Rubisco carboxylation rate ( V c,max ) at low (200 μmol⋅m –2 ⋅s –1 ) and high PPFD (1,000 μmol⋅m –2 ⋅s –1 ) were estimated from photosynthetic CO 2 response curves. Both QY m,inc (maximum QY on incident PPFD basis) and J at low PPFD were higher under red light than under blue and green light. Factoring in light absorption, QY m,abs (the maximum QY on absorbed PPFD basis) under green and red light were both higher than under blue light, indicating that the low QY m,inc under green light was due to lower absorptance, while absorbed blue photons were used inherently least efficiently. At high PPFD , the QY inc [gross CO 2 assimilation ( A g )/incident PPFD ] and J under red and green light were similar, and higher than under blue light, confirming our hypothesis. V c,max may not limit photosynthesis at a PPFD of 200 μmol m –2 s –1 and was largely unaffected by light spectrum at 1,000 μmol⋅m –2 ⋅s –1 . A g and J under different spectra were positively correlated, suggesting that the interactive effect between light spectrum and PPFD on photosynthesis was due to effects on J . No interaction between the three colors of light was detected. In summary, at low PPFD , green light had the lowest photosynthetic efficiency because of its low absorptance. Contrary, at high PPFD , QY inc under green light was among the highest, likely resulting from more uniform distribution of green light in leaves.
Introduction
The photosynthetic activity of light is wavelength dependent. Based on McCree’s work ( McCree, 1971 , 1972 ), photosynthetically active radiation is typically defined as light with a wavelength range from 400 to 700 nm. Light with a wavelength shorter than 400 nm or longer than 700 nm was considered as unimportant for photosynthesis, due to its low quantum yield of CO 2 assimilation, when applied as a single waveband ( Figure 1 ). Within the 400–700 nm range, McCree (1971) showed that light in the red region (600–700 nm) resulted in the highest quantum yield of CO 2 assimilation of plants. Light in the green region (500–600 nm) generally resulted in a slightly higher quantum yield than light in the blue region (400–500 nm) ( Figure 1 ; McCree, 1971 ). The low absorptance of green light is partly responsible for its low quantum yield of CO 2 assimilation. Within the visible spectrum, green leaves have the highest absorptance in the blue region, followed by red. Green light is least absorbed by green leaves, which gives leaves their green appearance ( McCree, 1971 ; Zhen et al., 2019 ).
Figure 1. The normalized action spectrum of the maximum quantum yield of CO 2 assimilation for narrow wavebands of light from ultra-violet to far-red wavelengths ( McCree, 1971 ). Redrawn using data from Sager et al. (1988).
Since red and blue light are absorbed more strongly by photosynthetic pigments than green light, they are predominantly absorbed by the top few cell layers, while green light can penetrate deeper into leaf tissues ( Nishio, 2000 ; Vogelmann and Evans, 2002 ; Terashima et al., 2009 ; Brodersen and Vogelmann, 2010 ), thus giving it the potential to excite photosystems in deeper cell layers. Leaf photosynthesis may benefit from the more uniform light distribution throughout a leaf under green light. Absorption of photons by chloroplasts near the adaxial surface may induce heat dissipation of excess excitation energy in those chloroplasts, while chloroplasts deeper into the leaf receive little excitation energy ( Sun et al., 1998 ; Nishio, 2000 ). Blue and red photons, therefore, may be used less efficiently and are more likely to be dissipated as heat than green photons.
The misconception that red and blue light are used more efficiently by plants than green light still occasionally appears ( Singh et al., 2015 ), often citing McCree’s action spectrum or the poor absorption of green light by chlorophyll extracts. The limitations of McCree’s action spectrum were explained in his original paper: the quantum yield was measured under low photosynthetic photon flux density ( PPFD ), using narrow waveband light, and expressed on an incident light basis ( McCree, 1971 ), but these limitations are sometimes ignored. The importance of green light for photosynthesis has been well established in more recent studies ( Sun et al., 1998 ; Nishio, 2000 ; Terashima et al., 2009 ; Hogewoning et al., 2012 ; Smith et al., 2017 ).
From those studies, one trend has emerged that has not received much attention: there is an interactive effect of light quality and intensity on photosynthesis ( Sun et al., 1998 ; Evans and Vogelmann, 2003 ; Terashima et al., 2009 ). At low PPFD , green light has the lowest QY inc (quantum yield of CO 2 assimilation on incident light basis) because of its low absorptance; at high PPFD , on the other hand, red and blue light have a lower QY inc than green light, because of their high absorptance by photosynthetic pigments, which shifts much of the light absorption closer to the upper leaf surface. This reduces both the quantum yield of CO 2 assimilation in cells in the upper part of a leaf and light availability in the bottom part of a leaf.
The interactive effect between light quality and intensity was illustrated in an elegant study that quantified the differential quantum yield, or the increase in leaf CO 2 assimilation per unit of additional light ( Terashima et al., 2009 ). The differential quantum yield was measured by adding red or green light to a background illumination of white light of different intensities. At low background white light levels, the differential quantum yield of red light was higher than that of green light, due to the low absorptance of green light. But as the background light level increased, the differential quantum yield of green light decreased more slowly than that of red light, and was eventually higher than that of red light ( Terashima et al., 2009 ). The red light was absorbed efficiently by the chloroplasts in the upper part of leaves. With a high background level of white light, those chloroplasts already received a large amount of excitation energy from white light and up-regulated non-photochemical quenching (NPQ) to dissipate excess excitation energy as heat, causing the additional red light to be used inefficiently. Green light, on the other hand, was able to reach the chloroplasts deeper in the mesophyll and excited those chloroplasts that received relatively little excitation energy from white light. Therefore, with high background white light intensity, additional green light increased leaf photosynthesis more efficiently than red light ( Terashima et al., 2009 ).
In this paper, we present a comprehensive study to explore potential interactive effect of light intensity and light quality on C 3 photosynthesis and underlying processes. We quantified the photosynthetic response of plants to blue, green, and red light over a wide PPFD range to better describe how light intensity and waveband interact. In addition, we examined potential interactions among blue, green, and red light, using light with different ratios and intensities of the three narrow waveband lights. To get a better understanding of the biochemical reasons for the effects of light spectrum and intensity on CO 2 assimilation, we constructed assimilation – internal leaf CO 2 ( C i ) response curves ( A/C i curves) under blue, green, and red light, as well as combinations of the three narrow waveband lights at both high and low PPFD . We hypothesized that effects of different light spectra would be reflected in the electron transport rate ( J ) required to regenerate consumed ribulose 1,5-bisphosphate (RuBP), rather than the maximum carboxylation rate of ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) ( V c,max ).
Materials and Methods
Plant material.
Lettuce “Green Towers” plants were grown from seed in 1.7 L round pots filled with soilless substrate (Fafard 4P Mix, Sun Gro Horticulture, Agawam, MA, United States). The plants were grown in a growth chamber (E15, Conviron, Winnipeg, Manitoba, Canada) at 23.2 ± 0.8°C (mean ± SD), under white fluorescent light with a 14-hr photoperiod, vapor pressure deficit (VPD) of 1.20 ± 0.43 kPa and a PPFD of 200–230 μmol⋅m –2 ⋅s –1 at the floor level, and ambient CO 2 concentration. Plants were sub-irrigated when necessary with a nutrient solution containing 100 mg⋅L –1 N, made with a complete, water-soluble fertilizer (Peter’s Excel 15-5-15 Cal-Mag fertilizer, Everris, Marysville, OH, United States).
Leaf Absorptance, Transmittance, and Reflectance
Leaf absorptance was determined using a method similar to that of Zhen et al. (2019) . Three plants were randomly selected. A newly expanded leaf from each plant was illuminated with a broad-spectrum halogen bulb (70W; Sylvania, Wilmington, MA, United States) for leaf transmittance measurement. Transmittance was measured with a spectroradiometer (SS-110, Apogee, Logan, UT, United States). The halogen light spectrum was taken as reference measurement with the spectroradiometer placed directly under the halogen bulb in a dark room. Then, a lettuce leaf was placed between the halogen bulb and spectroradiometer, with its adaxial side facing the halogen bulb and transmitted light was measured. Leaf transmittance was then calculated on 1 nm resolution. Light reflectance of the leaves was measured using a spectrometer with a leaf clip (UniSpec, PP systems, Amesbury, MA, United States). Light absorptance was calculated as 1− r e f l e c t a n c e − t r a n s m i t t a n c e . We verified that this method results in similar absorptance spectra as the use of an integrating sphere. Absorptance of each of the nine light spectra used in this study were calculated from the overall leaf absorptance spectrum and the spectra of the red, green, and blue LEDs.
Leaf Photosynthesis Measurements
All gas exchange measurements were made with a leaf gas exchange system (CIRAS-3, PP Systems). Light was provided by the LEDs built into the chlorophyll fluorescence module (CFM-3, PP Systems). This module has dimmable LED arrays of different colors, with peaks at 653 nm [red, full width at half maximum (FWHM) of 17 nm], 523 nm (green, FWHM of 36 nm), and 446 nm (blue, FWHM of 16 nm). Nine different combinations of red, green, and blue light were used in this study ( Table 1 ). Throughout the measurements, the environmental conditions inside the cuvette were controlled by the leaf gas exchange system. Leaf temperature was 23.0 ± 0.1°C, CO 2 concentration was 400.5 ± 4.1 μmol⋅mol –1 , and the VPD of air in the leaf cuvette was 1.8 ± 0.3 kPa (mean ± SD).
Table 1. List of light spectrum abbreviations and their spectral composition.
Photosynthesis – Light Response Curves
To explore photosynthetic efficiency of light with different spectra, we constructed light response curves for lettuce plants using each light spectrum. Lettuce plants were exposed to 10 PPFD levels ranging from 30 to 1,300 μmol⋅m –2 ⋅s –1 (30, 60, 90, 120, 200, 350, 500, 700, 1,000, and 1,300 μmol⋅m –2 ⋅s –1 ) in ascending orders for light response curves. Photosynthetic measurements were taken on 40–66 days old lettuce plants. Lettuce plants were taken out of the growth chamber and dark-adapted for 30 min. Starting from the lowest PPFD , one newly expanded leaf was exposed to all nine spectra. Net CO 2 assimilation rate ( A n ) of the leaf was measured using the leaf gas exchange system. Under each light spectrum, three A n readings were recorded at 10 s intervals after readings were stable (about 4–20 min depending on PPFD after changing PPFD and spectrum). The three A n readings were averaged for analysis. After A n measurements under all nine light spectra were taken, the leaf was exposed to the next PPFD level and A n measurements were taken with the light spectra in the same order, until measurements were completed at all PPFD levels. Throughout the light response curves, C i decreased with increasing PPFD , from 396 ± 10 μmol⋅mol –1 at a PPFD of 30 μmol⋅m –2 ⋅s –1 to 242 ± 44 μmol⋅mol –1 at a PPFD of 1,300 μmol⋅m –2 ⋅s –1 . To account for the potential effect of plants and the order of the spectra on assimilation rates, the order of the different spectra was re-randomized for each light response curve, using a Latin square design with plant and spectrum as the blocking factors. Data were collected on nine different plants.
Regression curves (exponential rise to maximum) were fitted to the data for each light spectrum and replication (plant):
where R d is the dark respiration rate, QY m,inc is the maximum quantum yield of CO 2 assimilation (initial slope of light response curve, mol of CO 2 fixed per mol of incident photons) and A g,max is the light-saturated gross assimilation rate. The A n,max is the light-saturated net assimilation rate and was calculated as A n , m a x = A g , m a x - R d . The maximum quantum yield of CO 2 assimilation was also calculated on absorbed light basis as Q Y m , a b s = Q Y m , i n c l i g h t a b s o r p t a n c e .
The instantaneous quantum yield of CO 2 assimilation based on incident PPFD ( QY inc ) was calculated as A g P P F D for each PPFD at which A n was measured, where the gross CO 2 assimilation rate ( A g ) was calculated as A g = A n + R d . To account for differences in absorptance among the different light spectra, the quantum yield of CO 2 assimilation was also calculated based on absorbed light base, as Q Y a b s = A g P P F D × l i g h t a b s o r p t a n c e , where light absorptance is the absorptance of lettuce leaves for each specific light spectrum. The differential QY , the increase in assimilation rate per unit of additional incident PPFD , was calculated as the derivative of Eq. 1:
Photosynthesis – Internal CO 2 Response ( A/C i ) Curves
To explore the underlying physiological mechanisms of assimilation responses to different light spectra, we constructed A/C i curves. Typically, A/C i curves are collected under saturating PPFD . We collected A/C i curves at two PPFD s (200 and 1,000 μmol⋅m –2 ⋅s –1 ) to explore interactive effects of light spectrum and PPFD on the assimilation rate. At a PPFD of 200 μmol⋅m –2 ⋅s –1 , red light has the highest A n and green light the lowest A n , while at PPFD of 1,000 μmol⋅m –2 ⋅s –1 , red and green light resulted in the highest A n and blue light in the lowest A n .
We used the rapid A/C i response (RACiR) technique that greatly accelerates the process of constructing A/C i curves ( Stinziano et al., 2017 ). We used a Latin square design, similar to the light response curves. A/C i curves were measured under the same nine spectra used for the light response curves. Nine lettuce plants were used as replicates. For each A/C i curve, CO 2 concentration in the leaf cuvette started from 0 μmol⋅mol –1 , steadily ramping to 1,200 μmol⋅mol –1 over 6 min. A reference measurement was also taken at the beginning of each replication with an empty cuvette to correct for the reaction time of the leaf gas exchange system. Post-ramp data processing was used to calculate the real A and C i with the spreadsheet provided by PP systems, which yielded the actual A/C i curves with C i range of about 100–950 μmol mol –1 . Throughout the data collection, leaf temperature was 24.4 ± 1.3°C and VPD in the cuvette was 1.4 ± 0.2 kPa.
Curve fitting for A/C i curves was done by minimizing the residual sum of squares, following the protocol developed by Sharkey et al. (2007) . Among our nine replicates, four plants did not show clear Rubisco limitations at low PPFD and for those plants Rubisco limitation ( V c,ma x ) was not included in the model ( Sharkey et al., 2007 ). We therefore report V c,max values for high PPFD only. The J was determined for all light spectra at both PPFD s. We therefore report V c,max was determined for all light spectra only at high PPFD . The quantum yield of electron transport [ QY(J) ] was calculated on both incident and absorbed PPFD basis as Q Y ( J ) i n c = J P P F D and Q Y ( J ) a b s = Q Y ( J ) i n c l i g h t a b s o r p t a n c e , respectively. We did not estimate triose phosphate utilization, because the A/C i curves often did not show a clear plateau.
Data Analysis
The QY m,inc , QY m,abs , and A g,max were analyzed with ANOVA to determine the effects of light spectrum using SAS (SAS University Edition; SAS Institute, Cary, NC, United States). A n , QY inc , and QY abs at each PPFD level and V c,max and J estimated from A/C i curves were similarly analyzed with ANOVA using SAS. A n at different PPFD levels were analyzed with regression analysis to detect interactive effect of blue, green, and red light on leaf assimilation rates using the fractions of red, blue, and green light as explanatory variables (JMP Pro 15, SAS Institute).
Leaf Absorptance
A representative spectrum of light absorptance, reflectance and transmittance of a newly fully expanded lettuce leaf is shown in Figure 2 . In the blue region, 400–500 nm, the absorptance by “Green Towers” lettuce leaves was high and fairly constant, averaging 91.6%. The leaf absorptance decreased as the wavelength increased from 500 to 551 nm where the absorptance minimum was 69.8%. Absorptance increased again at longer wavelengths, with a second peak at 666 nm (92.6%). Above 675 nm, the absorptance decreased steadily to <5% at 747 nm ( Figure 2 ). The absorptance spectrum of our lettuce leaves is similar to what McCree (1971) obtained for growth chamber-grown lettuce, with the exception of slightly higher absorptance in the green part of the spectrum in our lettuce plants. Using this spectrum, the absorptance of the blue, green, and red LED lights were calculated to be 93.2 ± 1.0%, 81.1 ± 1.9% and 91.6 ± 1.1%, respectively. Absorptance of all nine spectra was calculated based on their ratios of red, green, and blue light ( Table 2 ).
Figure 2. Light absorptance, reflectance, and transmittance spectrum of a newly fully expanded “Green Towers” lettuce leaf.
Table 2. Light absorptance and transmittance of new fully expanded “Green towers” lettuce leaves under nine light spectra.
Light Quality and Intensity Effects on Photosynthetic Parameters
Light response curves of lettuce under all nine spectra are shown in Figure 3 , with regression coefficients in Supplementary Table 1 . It is worth noting that a few plants showed photoinhibition under 100B (decrease in A n with PPFD > 1,000 μmol⋅m –2 ⋅s –1 ). Those data were excluded in curve fitting for light response curves to better estimate asymptotes. Photoinhibition was not observed under other spectra.
Figure 3. Net assimilation ( A n ) – light response curves of “Green Towers” lettuce under nine light spectra. Error bars represent the standard deviation ( n = 9). Inserts show A n against PPFD of 30-90 μmol⋅m –2 ⋅s –1 s to better show the initial slopes of curves. The composition of the nine light spectra is shown in Table 1 . The light spectra in the graphs are (A) 100B, 100G, and 100R; (B) 100B, 80B20G, 20B80G, and 100G; (C) 100G, 80G20R, 20G80R, and 100R; and (D) 20B80R, 16B20G64R, and 100G.
The QY m,inc of lettuce plants was 22 and 27% higher under red light (74.3 mmol⋅mol –1 ) than under either 100G (60.8 mmol⋅mol –1 ) or 100B (58.4 mmol⋅mol –1 ), respectively ( Figure 4A and Supplementary Table 1 ). Spectra with a high fraction of red light (64% or more) resulted in a high QY m,inc ( Figure 4A ), while 80G20R resulted in an intermediate QY m,inc ( Figure 4A ). To determine whether differences in QY m,inc were due to differences in absorptance or in the ability of plants to use the absorbed photons for CO 2 assimilation, we also calculated QY m,abs . On an absorbed light basis, 100B light still resulted in the lowest QY m,abs (62.7 mmol⋅mol –1 ) and red light resulted in the highest QY m,abs (81.1 mmol⋅mol –1 ) among narrow waveband lights ( Figure 4B ). Green light resulted in a QY m,abs (74.9 mmol⋅mol –1 ) similar to that under red light, but significantly higher than that of blue light ( Figure 4B ). We did not find any interactions (synergism or antagonism) between lights of different colors, with all physiological responses under mixed spectra being similar to the weighted average of responses under single colors. Thus, for the rest of the results we focus on the three narrow waveband spectra.
Figure 4. Maximum quantum yield of CO 2 assimilation of “Green Towers” lettuce based on incident ( QY m,inc ) (A) and absorbed light ( QY m,abs ) (B) under nine different light spectra. Values are calculated as the initial slope of the light response curves of corresponding light spectra (see Figure 3 ). Bars with the same letter are not statistically different ( p ≤ 0.05). Error bars represent the standard deviation ( n = 9). The composition of the nine light spectra is shown in Table 1 .
Among the three narrow waveband lights, 100G resulted in the highest A g,max (20.0 μmol⋅m –2 ⋅s –1 ), followed by red (18.9 μmol⋅m –2 ⋅s –1 ), and blue light (17.0 μmol⋅m –2 ⋅s –1 ) ( Figure 5 and Supplementary Table 1 ). As with QY m ,inc and QY m,abs , combining two or three colors of light resulted in an A g,max similar to the weighted averages of individual light colors.
Figure 5. Maximum gross assimilation rate ( A g,max ) of “Green Towers” lettuce under different light spectra, calculated from the light response curves. Bars with the same letter are not statistically different ( p ≤ 0.05). Error bars represent standard deviation ( n = 9). The composition of the nine light spectra is shown in Table 1 .
QY inc initially increased with increasing PPFD and peaked at 90–200 μmol⋅m –2 ⋅s –1 , then decreased at higher PPFDs ( Figure 6A ). The QY inc under 100R was higher than under either green or blue light at low PPFD (≤300 μmol⋅m –2 ⋅s –1 ). Although 100G resulted in lower QY inc than 100B at low PPFD (≤300 μmol⋅m –2 ⋅s –1 ), the decrease in QY inc under 100G with increasing PPFD was slower than that with 100B or 100R. Above 500 μmol m –2 s –1 , the QY inc with 100G was comparable to the QY inc with 100R, and higher than with 100B ( Figure 6A ). The QY abs with 100R was higher than that with either 100G or 100B at PPFDs from 60 to 120 μmol⋅m –2 ⋅s –1 ( p < 0.05). The QY abs with 100G was similar to 100B at low PPFD , but decreased slower than that with either 100R or 100B as PPFD increased. At PPFD ≥ 500 μmol⋅m –2 ⋅s –1 , QY abs was lowest under 100B among the three monochromatic lights ( p < 0.05) ( Figure 6B ).
Figure 6. The quantum yield of CO 2 assimilation of “Green Towers” lettuce as a function of incident ( QY inc ) (A) and absorbed PPFD ( QY abs ) (B) under blue, green, and red LED light. Error bars represent the standard deviation ( n = 9).
The differential QY , which quantifies the increase in CO 2 assimilation per unit of additional PPFD , decreased with increasing PPFD . The differential QY with 100R was higher than those with 100B and 100G at low PPFD . At a PPFD of 30 μmol⋅m –2 ⋅s –1 , the differential QY was 70.5 mmol⋅mol –1 for 100R, 59.4 mmol⋅mol –1 for 100G, and 55.8 mmol⋅mol –1 for 100B ( Figure 7 ). However, the differential QY with 100R decreased rapidly with increasing PPFD and was lower than the differential QY with 100G at high PPFD ( Figure 7 ). At high PPFD , the differential QY with 100G was highest among three monochromatic light ( Figure 7 ). For instance, at a PPFD of 1,300 μmol⋅m –2 ⋅s –1 , the differential QY with 100G was 1.09 mmol⋅mol –1 , while those with 100B and 100R were 0.64 mmol⋅mol –1 and 0.46 mmol⋅mol –1 , respectively ( Figure 7 ).
Figure 7. The differential quantum yield of CO 2 assimilation ( differential QY ) of “Green Towers” lettuce under blue, green, and red LED light as a function of the PPFD . The differential QY is the increase in net assimilation per unit additional PPFD and was calculated as the first derivate of the light response curves ( Figure 3 ). The insert shows the differential quantum yield plotted at PPFDs of 1,000–1,300 μmol m –2 s –1 s to better show differences at high PPFD (note the different y -axis scale).
Effect of Light Spectrum and Intensity on J and V c,max
J of lettuce leaves at low PPFD was lowest under 100G (47.4 μmol⋅m –2 ⋅s –1 ), followed by 100B (56.1 μmol⋅m –2 ⋅s –1 ), and highest under 100R (64.1 μmol⋅m –2 ⋅s –1 ) ( Figure 8A ). At high PPFD , on the other hand, J of leaves exposed to 100G (115.3 μmol⋅m –2 ⋅s –1 ) and 100R (112.1 μmol⋅m –2 ⋅s –1 ) were among the highest, while J of leaves under 100B was the lowest (97.0 μmol⋅m –2 ⋅s –1 ) ( Figure 8A ). At high PPFD , V c,max of leaves under blue light (59.3 μmol⋅m –2 ⋅s –1 ) was lower than V c,max of leaves under 16B20G64R light (72.1 μmol⋅m –2 ⋅s –1 ), but none of the other treatments differed significantly ( Figure 8 ). When PPFD increased from 200 to 1,000 μmol⋅m –2 ⋅s –1 , J under green light increased by 143%, while J under blue and red light increased by 73% and 75%, respectively ( Figure 8A ). J and V c,max at high PPFD were strongly correlated ( R 2 = 0.82) ( Supplementary Figure 3 ).
Figure 8. Electron transport rate ( J ) at PPFD s of 200 (left bars) and 1,000 μmol m –2 s –1 (right bars) (A) and maximum Rubisco carboxylation rate ( V c,max ) at a PPFD of 1,000 μmol m –2 s –1 (B) of “Green Towers” lettuce, as estimated from A/C i curves under different light spectra. Bars with the same letter are not statistically different ( p ≤ 0.05). Error bars represent the standard deviation ( n = 9). The light composition of the nine light spectra is shown in Table 1 .
Interactive Effect of Light Spectrum and PPFD on Photosynthesis
There was an interactive effect of light spectrum and PPFD on photosynthetic properties of lettuce. Under low light conditions (≤200 μmol⋅m –2 ⋅s –1 ), the QY inc of lettuce leaves under green light was lowest among blue, green, and red light ( Figure 6A ), due to its lower absorptance by lettuce leaves. After accounting for absorptance, green photons were used at similar efficiency as blue photons, while red photons were used most efficiently ( Figure 6B ). The QY m,abs under green and red light were higher than under blue light ( Figure 4B ). At high PPFD , green and red light had similar quantum yield, higher than that of blue light, both on an absorbed and incident light basis ( Figure 6A ). Multiple factors contributed to the interactive effect of light spectrum and PPFD on quantum yield and photosynthesis.
Light Absorptance and Non-Photosynthetic Pigments Determine Assimilation at Low PPFD
QY m,inc with blue and green light was lower than with red light ( Figure 4A ), consistent with McCree’s action spectrum ( McCree, 1971 ). But when taking leaf absorptance into account, QY m,abs was similar under green and red light and lower under blue light ( Figure 4B ). Similarly, at low PPFD (≤200 μmol⋅m –2 ⋅s –1 ), QY inc of lettuce leaves was highest under red, intermediate under blue, and lowest under green light. When accounting for leaf absorptance, QY abs under red light remained highest and QY abs under both green and blue light were similar at low PPFD ( Figure 6A ). Consistent with our data, previous studies also documented that, once absorbed, green light can drive photosynthesis efficiently at low PPFD ( Balegh and Biddulph, 1970 ; McCree, 1971 ; Evans, 1987 ; Sun et al., 1998 ; Nishio, 2000 ; Terashima et al., 2009 ; Hogewoning et al., 2012 ; Vogelmann and Gorton, 2014 ). For example, the QY m,abs of spinach ( Spinacia oleracea ) and cabbage ( Brassica oleracea L. ) was highest under red light, followed by that under green light and lowest with blue light. But on incident light basis, QY m,inc of under green light was lower than under red or blue light ( Sun et al., 1998 ).
Both our data ( Figure 4B ) and those of Sun et al. (1998) show that QY m,abs with blue light is lower than that with red and green light, indicating that blue light is used intrinsically less efficiently by lettuce. Blue light, and, to a lesser extent, green light is absorbed not just by chlorophyll, but also by flavonoids and carotenoids ( Sun et al., 1998 ). Those pigments can divert energy away from photochemistry and thus reduce the QY abs under blue light. Flavonoids (e.g., anthocyanins) are primarily located in the vacuole and cannot transfer absorbed light energy to photosynthetic pigments ( Sun et al., 1998 ). Likewise, free carotenoids do not contribute to photochemistry ( Hogewoning et al., 2012 ). Carotenoids in light-harvesting antennae and reaction centers channel light energy to photochemistry, but with lower transfer efficiency than chlorophylls ( Croce et al., 2001 ; de Weerd et al., 2003a , b ; Wientjes et al., 2011 ; Hogewoning et al., 2012 ). Therefore, absorption of blue light by flavonoids and carotenoids reduces the quantum yield of CO 2 assimilation. Thus, even with the high absorptance of blue light by green leaves, QY m,abs of leaves under blue light was the lowest among the three monochromatic lights ( Figure 4B ). It is likely that the lower QY abs under green light than that under red light was also due to absorption of green light by carotenoids and flavonoids ( Hogewoning et al., 2012 ). At high PPFD , absorption of blue light by flavonoids and carotenoids still occurs, but this is less of a limiting factor for photosynthesis, since light availability is not limiting under high PPFD .
Light Dependence of Respiration and Rubisco Activity May Reduce the Quantum Yield at Low PPFD
At PPFD s below 200 μmol⋅m –2 ⋅s –1 , the QY inc and QY abs of lettuce showed an unexpected pattern in response to PPFD ( Figure 6 ). Unlike the quantum yield of PSII, which decreases exponentially with increasing PPFD ( Weaver and van Iersel, 2019 ), QY inc and QY abs increased initially with increasing PPFD ( Figure 6 ). A similar pattern was previously observed by Craver et al. (2020) in petunia ( Petunia × hybrida ) seedlings. This pattern could result from light-dependent regulation of respiration ( Croce et al., 2001 ), alternative electron sinks such as nitrate reduction ( Skillman, 2008 ; Nunes-Nesi et al., 2010 ), or Rubisco activity ( Campbell and Ogren, 1992 ; Zhang and Portis, 1999 ). In our calculations, we assumed that the leaf respiration in the light was the same as R d . However, leaf respiration in the light is lower than in the dark, in a PPFD -dependent manner ( Brooks and Farquhar, 1985 ; Atkin et al., 1997 ), which can lead to overestimation of A g with increasing PPFD . When we accounted for this down-regulation of respiration, using the model by Müller et al. (2005) to correct A g , QY inc , and QY abs , we found that depression of respiration by light did not explain the initial increase in QY inc and QY abs we observed ( Supplementary Figure 4 ). Alternative electron sinks in the chloroplasts that are upregulated in response to light can explain the low QY inc , and QY abs at low PPFD , because they compete with the Calvin cycle for reducing power (ferredoxin/NADPH). Such processes include photorespiration ( Krall and Edwards, 1992 ), nitrate assimilation ( Nunes-Nesi et al., 2010 ), sulfate assimilation ( Takahashi et al., 2011 ) and the Mehler reaction ( Badger et al., 2000 ) and their effect on QY inc , and QY abs would be especially notable under low PPFD ( Supplementary Figure 5 ).
Upregulation of Rubisco activity by Rubisco activase in the light may also have contributed to the increase in QY inc and QY abs at low PPFD ( Campbell and Ogren, 1992 ; Zhang and Portis, 1999 ). In the dark, 2-carboxy-D-arabinitol-1-phosphate (CA1P) or RuBP binds strongly to the active sites of Rubisco, preventing carboxylation activity. In the light, Rubisco activase releases the inhibitory CA1P or RuBP from the catalytic site of Rubisco, in a light-dependent manner ( Campbell and Ogren, 1992 ; Zhang and Portis, 1999 ; Parry et al., 2008 ). At PPFD < 120 μmol⋅m –2 ⋅s –1 , low Rubisco activity may have limited photosynthesis.
Light Distribution Within Leaves Affects QY at High PPFD
Except for the initial increase at low PPFD , both QY inc and QY abs decreased with increasing PPFD . QY inc decreased slower under green than under red or blue light ( Figure 6A ). At a PPFD ≥ 500 μmol⋅m –2 ⋅s –1 , QY inc under green light was higher than that under blue light ( Figure 6A ). Accordingly, A n under blue light was lower than under green and red light at PPFD s above 500 μmol⋅m –2 ⋅s –1 ( Figure 3A ). The lower QY inc under blue light than under green and red light at high PPFD can be explained by disparities in the light distribution within leaves.
Blue and red light were strongly absorbed by lettuce leaves (93.2 and 91.6%, respectively), while green light was absorbed less (81.1%) ( Table 2 ). Similar low green absorptance was found in sunflower ( Helianthus annuus L.), snapdragon ( Antirrhínum majus L.) ( Brodersen and Vogelmann, 2010 ), and spinach ( Vogelmann and Han, 2000 ). In leaves of those species, absorption of red and blue light peaked in the upper 20% of leaves, and declined sharply further into the leaf. Absorption of red light decreased slower with increasing depth than that of blue light ( Vogelmann and Han, 2000 ; Brodersen and Vogelmann, 2010 ). Green light absorption peaked deeper into leaves, and was more evenly distributed throughout leaves, because of low absorption of green light by chlorophyll ( Vogelmann and Han, 2000 ; Brodersen and Vogelmann, 2010 ). The more even distribution of green light within leaves, as compared to red and blue light, can explain the interactive effects between PPFD and light spectrum on leaf photosynthesis. It was estimated that less than 10% of blue light traveled through the palisade mesophyll and reached the spongy mesophyll in spinach, while about 35% of green light and 25% of red light did so ( Vogelmann and Evans, 2002 ). It was also estimated that chlorophyll in the lowermost chloroplasts of spinach leaves absorbed about 10% of green and <2% of blue light, compared to chlorophyll in the uppermost chloroplasts ( Vogelmann and Evans, 2002 ; Terashima et al., 2009 ).
The more uniform green light distribution within leaves may be a key contributor to higher leaf level QY inc under high PPFD because less heat dissipation of excess light energy is needed ( Nishio, 2000 ; Terashima et al., 2009 ). Reaction centers near the adaxial leaf surface receive more excitation energy under blue, and to a lesser extent under red light, than under green light, because of the differences in absorptance. Consequently, under high intensity blue light, NPQ is up-regulated in the chloroplasts near the adaxial leaf surface to dissipate some of the excitation energy ( Sun et al., 1998 ; Nishio, 2000 ), lowering the QY inc under blue light. Since less green light is absorbed near the adaxial surface, less heat dissipation is required. When incident light increased from 150 to 600 μmol⋅m –2 ⋅s –1 , the fraction of whole leaf CO 2 assimilation that occurred in the top half of spinach leaves remained the same under green light (58%), but decreased from 87 to 73% under blue light. This indicates more upregulation of heat dissipation in the top of the leaves under blue, than under green light ( Evans and Vogelmann, 2003 ). On the other hand, the bottom half of the leaves can still utilize the available light with relatively high QY inc , since the amount of light reaching the bottom half is relatively low, even under high PPFD ( Nishio, 2000 ). By channeling more light to the under-utilized bottom part of leaves, leaves could achieve higher QY inc even under high intensity green light. In our study, high QY inc under green light and low QY inc under blue light at high PPFD ( Figure 6 ) can be thus explained by the large disparities in the light environment in chloroplasts from the adaxial to the abaxial side of leaves due to differences in leaf absorptance. Similarly, differential QY of lettuce leaves was highest under green light and lower under blue and red light at high PPFD (>300 μmol⋅m –2 ⋅s –1 ) ( Figure 7 ), also potentially because of the more uniform distribution of green light and the uneven distribution of blue and red light in leaves.
Along the same line, A n of lettuce leaves was the lowest under blue light at PPFD > 500 μmol⋅m –2 ⋅s –1 ( Figure 3 ). Also, A n of lettuce leaves approached light saturation at lower PPFD s under blue and red light, than under green light ( Figure 3A ). Under blue, green, and red light, lettuce leaves reached 95% of A n,max at PPFD s of 954, 1,110 and 856 μmol⋅m –2 ⋅s –1 , respectively. This can be seen more clearly in the differential QY at high PPFD ( Figure 7 ). At a PPFD of 1,300 μmol⋅m –2 ⋅s –1 , green light had a differential QY of 1.09 mmol⋅mol –1 , while that of red and blue light was only 0.46 and 0.69 mmol⋅mol –1 , respectively ( Figure 7 ). Green light also resulted in a higher A g,max (22.9 μmol⋅m –2 ⋅s –1 ) than red and blue light (21.8 and 19.3 μmol⋅m –2 ⋅s –1 , respectively) ( Figure 5 ). As discussed before, the high A g,max under green light resulted from the more uniform light distribution under green light, allowing deeper cell layers to photosynthesize more. Previous research similarly found that at high PPFD (>500 μmol⋅m –2 ⋅s –1 ), A n of both spinach and cabbage were lower under blue light than under white, red and green light ( Sun et al., 1998 ). Overall, under high PPFD , the differences in light distribution throughout a leaf are important to quantum yield and assimilation rate, since it affects NPQ up-regulation ( Sun et al., 1998 ; Nishio, 2000 ). However, light distribution within a leaf is less important at low than at high PPFD , because upregulation of NPQ increases with increasing PPFD ( Zhen and van Iersel, 2017 ).
Light Spectrum Affects J , but Not V c,max
We examined the effect of light quality and intensity on J and V c,max ( Figure 8 ). For the light-dependent reactions, the interactive effect between light spectra and PPFD found for CO 2 assimilation and quantum yield was also observed for J ( Figure 8A ). At low PPFD (200 μmol⋅m –2 ⋅s –1 ), green light resulted in the lowest J and red light in the highest J among single waveband spectra. But at a PPFD of 1,000 μmol⋅m –2 ⋅s –1 , red and green light resulted in the highest J and blue light in the lowest J ( Figure 8A ), similar to the differences in A g .
There was no clear evidence of Rubisco limitations to photosynthesis at a PPFD of 200 μmol⋅m –2 ⋅s –1 , so the rate of the light-dependent reactions likely limited photosynthesis. This is corroborated by the strong correlation between A g and J at a PPFD of 200 μmol⋅m –2 ⋅s –1 . Although Rubisco limitations to photosynthesis were observed at a PPFD of 1,000 μmol⋅m –2 ⋅s –1 , there were no meaningful differences in V c,max in response to light spectrum, in contrast to J ( Figure 8 ).
When PPFD increased 5×, from 200 to 1,000 μmol⋅m –2 ⋅s –1 , there was only a 1.7 to 2.4× increase in J , indicating a lower QY(J) inc at higher PPFD . This matches the lower QY inc and the asymptotic increase in A n in response to increasing PPFD ( Figure 3 ). The relative increase of J under green light (143%) was greater than that under both blue and red light (73 and 75%, respectively) as PPFD increased. This similarly can be attributed to a more uniform energy distribution of green light among reaction centers throughout a leaf and weaker upregulation of non-photochemical quenching with increasing green light intensity ( Sun et al., 1998 ; Nishio, 2000 ; Evans and Vogelmann, 2003 ), as discussed before.
There was a strong correlation between J and A g under the nine light spectra at both PPFD levels ( Figure 9A ). QY abs and QY(J) abs are similarly strongly correlated ( Figure 9B ). Unlike J , V c,max was largely unaffected by light spectra ( Figure 8B ) and was not correlated with A g (data not shown). There was, however, a strong correlation between J and V c,max at a PPFD of 1,000 μmol⋅m –2 ⋅s –1 ( R 2 = 0.82, Supplementary Figure 3 ), suggesting that J and V c,max are co-regulated. Similarly, Wullschleger (1993) noted a strong linear relationship between J and V c,max across 109 C 3 species. The ratio between J and V c,max in our study (1.5–2.0) similar to the ratio found by Wullschleger (1993) . These results suggest that the interactive effect of light spectra and PPFD resulted from effects on J , which is associated with light energy harvesting by reaction centers, rather than from V c,max .
Figure 9. The correlation between gross CO 2 assimilation rate ( A g ) estimated from light response curves and electron transport rate ( J ) estimated from A/C i curves (A) , and between the quantum yield of CO 2 assimilation ( QY abs ) and the quantum yield of electron transport on an absorbed light basis [ QY(J) abs ] (B) , under low PPFD (200 μmol m –2 s –1 ) and high PPFD (1,000 μmol m –2 s –1 ) under nine light spectra averaged over nine “Green Towers” lettuce plants. The color scheme representing the nine spectra is the same as Figure 8 .
No Interactive Effects Among Blue, Green, and Red Light
The Emerson enhancement effect describes a synergistic effect between lights of different wavebands (red and far-red) on photosynthesis ( Emerson, 1957 ). McCree (1971) attempted to account for interactions between light with different spectra when developing photosynthetic action spectra and applied low intensity monochromatic lights from 350 to 725 nm with white background light to plants. His results showed no interactive effect between those monochromatic lights and white light ( McCree, 1971 ). We tested different ratios of blue, green, and red light and different PPFD s, and similarly did not find any synergistic or antagonistic effect of different wavebands on any physiological parameters measured or calculated.
Importance of Interactions Between PPFD and Light Quality and Its Applications
The interactive effect between PPFD and light quality demonstrates a remarkable adaptation of plants to different light intensities. By not absorbing green light strongly, plants open up a “green window,” as Terashima et al. (2009) called it, to excite chloroplasts deeper into leaves, and thus facilitating CO 2 assimilation throughout the leaf. While red light resulted in relatively high QY inc , QY abs and A n at both high and low PPFD ( Figures 3 , 6 ), it is still mainly absorbed in the upper part of leaves ( Sun et al., 1998 ; Brodersen and Vogelmann, 2010 ). Green light can penetrate deeper into leaves ( Brodersen and Vogelmann, 2010 ) and help plants drive efficient CO 2 assimilation at high PPFD ( Figures 3 , 5 ).
Many early photosynthesis studies investigated the absorptance and action spectrum of photosynthesis of green algae, e.g., Haxo and Blinks (1950) or chlorophyll or chloroplasts extracts, e.g., Chen (1952) . Extrapolating light absorptance of green algae and suspension of chlorophyll or chloroplast to whole leaves from can lead to an underestimation of absorptance of green light by whole leaves and the belief that green light has little photosynthetic activity ( Moss and Loomis, 1952 ; Smith et al., 2017 ). Photosynthetic action spectra developed on whole leaves of higher plants, however, have long shown that green light effectively contributes to CO 2 assimilation, although with lower QY inc than red light ( Hoover, 1937 ; McCree, 1971 ; Inada, 1976 ; Evans, 1987 ). The importance of green light for photosynthesis was clearly established in more recent studies, emphasizing its role in more uniformly exciting all chloroplasts, which especially important under high PPFD ( Sun et al., 1998 ; Nishio, 2000 ; Terashima et al., 2009 ; Hogewoning et al., 2012 ; Smith et al., 2017 ). The idea that red and blue light are more efficient at driving photosynthesis, unfortunately, still lingers, e.g., Singh et al. (2015) .
Light-emitting diodes (LEDs) have received wide attention in recent years for use in controlled environment agriculture, as they now have superior efficacy over traditional lighting technologies ( Pattison et al., 2018 ). LEDs can have a narrow spectrum and great controllability. This provides unprecedented opportunities to fine tune light spectra and PPFD to manipulate crop growth and development. Blue and red LEDs have higher efficacy than white and green LEDs ( Kusuma et al., 2020 ). By coincidence, McCree’s action spectrum ( Figure 1 ; McCree, 1971 ) also has peaks in the red and blue region, although the peak in the blue region is substantially lower than the one in the red region. Therefore, red and blue LEDs are sometimes considered optimal for driving photosynthesis. This claim holds true only under low PPFD . Green light plays an important role in photosynthesis, as it helps plants to adapt to different light intensities. The wavelength-dependent absorptance of chlorophylls channels green light deeper into leaves, resulting in more uniform light absorption throughout leaves and providing excitation energy to cells further from the adaxial surface. Under high PPFD , this can increase leaf photosynthesis. Plant evolved under sunlight for hundreds of millions of years, and it seems likely that the relatively low absorptance of green light contributes to the overall photosynthetic efficiency of plants ( Nishio, 2000 ).
There was an interactive effect of light spectrum and PPFD on leaf photosynthesis. Under low PPFD , QY inc was lowest under green and highest under red light. The low QY inc under green light at low PPFD was due to low absorptance. In contrast, at high PPFD , green and red light achieved similar QY inc , higher than that of blue light. The strong absorption of blue light by chlorophyll creates a large light gradient from the top to the bottom of leaves. The large amount of excitation energy near the adaxial side of a leaf results in upregulation of nonphotochemical quenching, while chloroplasts near the bottom of a leaf receive little excitation energy under blue light. The more uniform distribution of green light absorption within leaves reduces the need for nonphotochemical quenching near the top of the leaf, while providing more excitation energy to cells near the bottom of the leaf. We also found that the interactive effect of light spectrum and PPFD on photosynthesis was a result of the light-dependent reactions; gross assimilation and J were strongly correlated. We detected no synergistic or antagonistic interactions between blue, green, and red light.
Data Availability Statement
The raw data supporting the conclusions of this article will be made available by the authors, without undue reservation.
Author Contributions
JL and MI designed the experiment, discussed the data, and revised the manuscript. JL performed the experiment, analyzed data, and prepared the first draft. Both authors contributed to the article and approved the submitted version.
This study was funded by the USDA-NIFA-SCRI award number 2018-51181-28365, project Lighting Approaches to Maximize Profits.
Conflict of Interest
The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.
Supplementary Material
The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fpls.2021.619987/full#supplementary-material
Supplementary Figure 1 | (Related to Figure 6 ) Quantum yield of CO 2 assimilation of “Green Towers” lettuce as a function of incident ( QY inc ) (A,C,E,G) and absorbed PPFD ( QY abs ) (B,D,F,H) under nine light spectra (see Table 1 ). Error bars represent standard deviation ( n = 9).
Supplementary Figure 2 | (Related to Figure 7 ) Differential quantum yield of CO 2 assimilation ( differential QY ) of “Green Towers” lettuce under nine light spectra as a function of the PPFD . Inserts show differential QY at PPFD s of 1,000–1,300 μmol⋅m –2 s –1 s to better show differences at high PPFD (note the different y -axis scale). The composition of the nine light spectra is shown in Table 1 . The light spectra in the graphs are (A) 100B, 100G and 100R; (B) 100B, 80B20G, 20B80G and 100G; (C) 100G, 80G20R, 20G80R and 100R; and (D) 20B80R, 16B20G64R and 100G.
Supplementary Figure 3 | (Related to Figure 6 ) The correlation between electron transport ( J ) and maximum Rubisco carboxylation rate ( V c,max ) of “Green Towers” lettuce estimated from A/C i curves under PPFD (1000 μmol m –2 s –1 ) under nine light spectra ( p < 0.001).
Supplementary Figure 4 | (Related to Figure 6 ) The comparison between QY inc before (A) and after (B) correcting for light-suppression of respiration under blue, green, and red LED light. Note that the initial increase in QY inc became more pronounced after correction of light suppressed respiration.
Supplementary Figure 5 | The comparison between QY abs before (A) and after (B) correcting for alternative electron sinks under blue, green, and red LED light. Assuming a simplified electron sink that diverts energy of 15 μmol m –2 s –1 of absorbed photons (an arbitrary value used for illustrative purposes only) away from the Calvin cycle under all PPFD s, the corrected QY abs was calculated based on remaining photons available to support Calvin cycle processes (B) . Note that the pattern of QY inc after correcting of alternative electron sink (B) is similar to quantum yield of PSII measured by chlorophyll fluorescence by Weaver and van Iersel (2019) .
Supplementary Table 1 | Dark respiration rate (R d ), maximum quantum yield of CO 2 assimilation (QY m,inc ) and maximum gross assimilation rate (A g,max ) of “Green towers” lettuce derived from the light response curves for nine different spectra using Eq. 1. The light response curves are shown in Figure 3 . *See light composition of nine lights presented here in Table 1 .
Abbreviations
PPFD , photosynthetic photon flux density; RuBP, ribulose 1,5-bisphosphate; Rubisco, ribulose-1,5-bisphosphate carboxylase/oxygenase; VPD, vapor pressure deficit; FWHM, full width at half maximum; A n , net CO 2 assimilation rate; R d , dark respiration rate; QY m,inc , maximum quantum yield of CO 2 assimilation; A g,max , light-saturated gross assimilation rate; QY m,abs , maximum quantum yield of CO 2 assimilation on absorbed light base; QY inc , quantum yield of CO 2 assimilation based on incident PPFD ; A g , gross CO 2 assimilation rate; QY abs , quantum yield of CO 2 assimilation on absorbed light base; QY , quantum yield of CO 2 assimilation; A/Ci curve, assimilation – internal leaf CO 2 response curve; RACiR, rapid A/C i response technique; V c,max , maximum rate of Rubisco carboxylation; J , rate of electron transport; CA1P, 2-carboxy- D -arabinitol-1-phosphate; NPQ, non-photochemical quenching.
Atkin, O. K., Westbeek, M., Cambridge, M. L., Lambers, H., and Pons, T. L. (1997). Leaf respiration in light and darkness (A comparison of slow- and fast-growing Poa species). Plant Physiol. 113, 961–965. doi: 10.1104/pp.113.3.961
PubMed Abstract | CrossRef Full Text | Google Scholar
Badger, M. R., von Caemmerer, S., Ruuska, S., and Nakano, H. (2000). Electron flow to oxygen in higher plants and algae: rates and control of direct photoreduction (Mehler reaction) and rubisco oxygenase. Philos. Trans. R. Soc. Lond. B Biol. Sci. 355, 1433–1446. doi: 10.1098/rstb.2000.0704
Balegh, S. E., and Biddulph, O. (1970). The photosynthetic action spectrum of the bean plant. Plant Physiol. 46, 1–5. doi: 10.1104/pp.46.1.1
Brodersen, C. R., and Vogelmann, T. C. (2010). Do changes in light direction affect absorption profiles in leaves? Funct. Plant Biol. 37, 403–412. doi: 10.1071/fp09262
CrossRef Full Text | Google Scholar
Brooks, A., and Farquhar, G. D. (1985). Effect of temperature on the CO2/O2 specificity of ribulose-1,5-bisphosphate carboxylase/oxygenase and the rate of respiration in the light. Planta 165, 397–406. doi: 10.1007/BF00392238
Campbell, W. J., and Ogren, W. L. (1992). Light activation of rubisco by rubisco activase and thylakoid membranes. Plant Cell Physiol. 33, 751–756. doi: 10.1093/oxfordjournals.pcp.a078314
Chen, S. L. (1952). The action spectrum for the photochemical evolution of oxygen by isolated chloroplasts. Plant Physiol. 27, 35–48. doi: 10.1104/pp.27.1.35
Craver, J. K., Nemali, K. S., and Lopez, R. G. (2020). Acclimation of growth and photosynthesis in Petunia seedlings exposed to high-intensity blue radiation. J. Am. Soc. Hortic. Sci. 145, 152–161. doi: 10.21273/jashs04799-19
Croce, R., Müller, M. G., Bassi, R., and Holzwarth, A. R. (2001). Carotenoid-to-chlorophyll energy transfer in recombinant major light-harvesting complex (LHCII) of higher plants. I. Femtosecond transient absorption measurements. Biophys. J. 80, 901–915. doi: 10.1016/S0006-3495(01)76069-9
de Weerd, F. L., Dekker, J. P., and van Grondelle, R. (2003a). Dynamics of β-carotene-to-chlorophyll singlet energy transfer in the core of photosystem II. J. Phys. Chem. B 107, 6214–6220. doi: 10.1021/jp027737q
de Weerd, F. L., Kennis, J. T., Dekker, J. P., and van Grondelle, R. (2003b). β-Carotene to chlorophyll singlet energy transfer in the photosystem I core of Synechococcus elongatus proceeds via the β-carotene S2 and S1 states. J. Phys. Chem. B 107, 5995–6002. doi: 10.1021/jp027758k
Emerson, R. (1957). Dependence of yield of photosynthesis in long-wave red on wavelength and intensity of supplementary light. Science 125, 746–746. doi: 10.1126/science.125.3251.746
Evans, J. (1987). The dependence of quantum yield on wavelength and growth irradiance. Funct. Plant Biol. 14, 69–79. doi: 10.1071/PP9870069
Evans, J., and Vogelmann, T. C. (2003). Profiles of 14C fixation through spinach leaves in relation to light absorption and photosynthetic capacity. Plant Cell Environ. 26, 547–560. doi: 10.1046/j.1365-3040.2003.00985.x
Haxo, F. T., and Blinks, L. (1950). Photosynthetic action spectra of marine algae. J. Gen. Physiol. 33, 389–422. doi: 10.1085/jgp.33.4.389
Hogewoning, S. W., Wientjes, E., Douwstra, P., Trouwborst, G., Van Ieperen, W., Croce, R., et al. (2012). Photosynthetic quantum yield dynamics: from photosystems to leaves. Plant Cell 24, 1921–1935. doi: 10.1105/tpc.112.097972
Hoover, W. H. (1937). The dependence of carbon dioxide assimilation in a higher plant on wave length of radiation. Smithson. Misc. Collect. 95, 1–13.
Google Scholar
Inada, K. (1976). Action spectra for photosynthesis in higher plants. Plant Cell Physiol. 17, 355–365. doi: 10.1093/oxfordjournals.pcp.a075288
Krall, J. P., and Edwards, G. E. (1992). Relationship between photosystem II activity and CO2 fixation in leaves. Physiol. Plant. 86, 180–187. doi: 10.1111/j.1399-3054.1992.tb01328.x
Kusuma, P., Pattison, P. M., and Bugbee, B. (2020). From physics to fixtures to food: current and potential LED efficacy. Hortic. Res. 7:56. doi: 10.1038/s41438-020-0283-7
McCree, K. J. (1971). The action spectrum, absorptance and quantum yield of photosynthesis in crop plants. Agric. Meteorol. 9, 191–216. doi: 10.1016/0002-1571(71)90022-7
McCree, K. J. (1972). Test of current definitions of photosynthetically active radiation against leaf photosynthesis data. Agric. Meteorol. 10, 443–453. doi: 10.1016/0002-1571(72)90045-3
Moss, R. A., and Loomis, W. E. (1952). Absorption spectra of leaves. I. the visible spectrum. Plant Physiol. 27, 370–391. doi: 10.1104/pp.27.2.370
Müller, J., Wernecke, P., and Diepenbrock, W. (2005). LEAFC3-N: a nitrogen-sensitive extension of the CO 2 and H 2 O gas exchange model LEAFC3 parameterised and tested for winter wheat (Triticum aestivum L.). Ecol. Modell. 183, 183–210. doi: 10.1016/j.ecolmodel.2004.07.025
Nishio, J. (2000). Why are higher plants green? Evolution of the higher plant photosynthetic pigment complement. Plant Cell Environ. 23, 539–548. doi: 10.1046/j.1365-3040.2000.00563.x
Nunes-Nesi, A., Fernie, A. R., and Stitt, M. (2010). Metabolic and signaling aspects underpinning the regulation of plant carbon nitrogen interactions. Mol. plant 3, 973–996.
Parry, M. A., Keys, A. J., Madgwick, P. J., Carmo-Silva, A. E., and Andralojc, P. J. (2008). Rubisco regulation: a role for inhibitors. J. Exp. Bot. 59, 1569–1580. doi: 10.1093/jxb/ern084
Pattison, P., Lee, K., Stober, K., and Yamada, M. (2018). Energy Savings Potential of SSL in Horticultural Applications. Washington, DC: U.S. Department of Energy, Office of Scientific and Technical Information.
Sharkey, T. D., Bernacchi, C. J., Farquhar, G. D., and Singsaas, E. L. (2007). Fitting photosynthetic carbon dioxide response curves for C3 leaves. Plant Cell Environ. 30, 1035–1040. doi: 10.1111/j.1365-3040.2007.01710.x
Singh, D., Basu, C., Meinhardt-Wollweber, M., and Roth, B. (2015). LEDs for energy efficient greenhouse lighting. Renew. Sustain. Energy Rev. 49, 139–147.
Skillman, J. B. (2008). Quantum yield variation across the three pathways of photosynthesis: not yet out of the dark. J. Exp. Bot. 59, 1647–1661. doi: 10.1093/jxb/ern029
Smith, H. L., McAusland, L., and Murchie, E. H. (2017). Don’t ignore the green light: exploring diverse roles in plant processes. J. Exp. Bot. 68, 2099–2110. doi: 10.1093/jxb/erx098
Stinziano, J. R., Morgan, P. B., Lynch, D. J., Saathoff, A. J., McDermitt, D. K., and Hanson, D. T. (2017). The rapid A–Ci response: photosynthesis in the phenomic era. Plant Cell Environ. 40, 1256–1262. doi: 10.1111/pce.12911
Sun, J., Nishio, J. N., and Vogelmann, T. C. (1998). Green light drives CO2 fixation deep within leaves. Plant Cell Physiol. 39, 1020–1026. doi: 10.1093/oxfordjournals.pcp.a029298
Takahashi, H., Kopriva, S., Giordano, M., Saito, K., and Hell, R. (2011). Sulfur assimilation in photosynthetic organisms: molecular functions and regulations of transporters and assimilatory enzymes. Annu. Rev. Plant Biol. 62, 157–184. doi: 10.1146/annurev-arplant-042110-103921
Terashima, I., Fujita, T., Inoue, T., Chow, W. S., and Oguchi, R. (2009). Green light drives leaf photosynthesis more efficiently than red light in strong white light: revisiting the enigmatic question of why leaves are green. Plant cell physiol. 50, 684–697. doi: 10.1093/pcp/pcp034
Vogelmann, T., and Han, T. (2000). Measurement of gradients of absorbed light in spinach leaves from chlorophyll fluorescence profiles. Plant Cell Environ. 23, 1303–1311. doi: 10.1046/j.1365-3040.2000.00649.x
Vogelmann, T. C., and Evans, J. (2002). Profiles of light absorption and chlorophyll within spinach leaves from chlorophyll fluorescence. Plant Cell Environ. 25, 1313–1323. doi: 10.1046/j.1365-3040.2002.00910.x
Vogelmann, T. C., and Gorton, H. L. (2014). “Leaf: light capture in the photosynthetic organ,” in The Structural Basis of Biological Energy Generation , ed. M. F. Hohmann-Marriott, (Dordrecht: Springer Netherlands), 363–377.
Weaver, G., and van Iersel, M. W. (2019). Photochemical characterization of greenhouse-grown Lettuce ( Lactuca sativa L. ‘Green Towers’) with applications for supplemental lighting control. HortScience 54, 317–322. doi: 10.21273/hortsci13553-18
Wientjes, E., van Stokkum, I. H., van Amerongen, H., and Croce, R. (2011). The role of the individual Lhcas in photosystem I excitation energy trapping. Biophys. J. 101, 745–754. doi: 10.1016/j.bpj.2011.06.045
Wullschleger, S. D. (1993). Biochemical limitations to carbon assimilation in C3 Plants—a retrospective analysis of the A/Ci curves from 109 species. J. Exp. Bot. 44, 907–920. doi: 10.1093/jxb/44.5.907
Zhang, N., and Portis, A. R. (1999). Mechanism of light regulation of rubisco: a specific role for the larger rubisco activase isoform involving reductive activation by thioredoxin-f. Proc. Natl. Acad. Sci. U.S.A. 96, 9438–9443. doi: 10.1073/pnas.96.16.9438
Zhen, S., Haidekker, M., and van Iersel, M. W. (2019). Far−red light enhances photochemical efficiency in a wavelength−dependent manner. Physiol. Plant. 167, 21–33. doi: 10.1111/ppl.12834
Zhen, S., and van Iersel, M. W. (2017). Photochemical acclimation of three contrasting species to different light levels: implications for optimizing supplemental lighting. J. Am. Soc. Hortic. Sci. 142, 346–354. doi: 10.21273/jashs04188-17
Keywords : photosynthesis, quantum yield of CO 2 assimilation, light spectrum, photosynthetic photon flux density, electron transport, V c, max , light intensity, light quality
Citation: Liu J and van Iersel MW (2021) Photosynthetic Physiology of Blue, Green, and Red Light: Light Intensity Effects and Underlying Mechanisms. Front. Plant Sci. 12:619987. doi: 10.3389/fpls.2021.619987
Received: 21 October 2020; Accepted: 11 February 2021; Published: 05 March 2021.
Reviewed by:
Copyright © 2021 Liu and van Iersel. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY) . The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.
*Correspondence: Jun Liu, [email protected]
Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.
Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.
- View all journals
- Explore content
- About the journal
- Publish with us
- Sign up for alerts
- Open access
- Published: 17 October 2024
Potential for photosynthesis on Mars within snow and ice
- Aditya R. Khuller ORCID: orcid.org/0000-0002-9866-781X 1 ,
- Stephen G. Warren ORCID: orcid.org/0000-0001-5429-7100 2 ,
- Philip R. Christensen 3 &
- Gary D. Clow 4
Communications Earth & Environment volume 5 , Article number: 583 ( 2024 ) Cite this article
Metrics details
- Astrobiology
- Cryospheric science
- Inner planets
On Earth, solar radiation can transmit down to multiple metres within ice, depending on its optical properties. Organisms within ice can harness energy from photosynthetically active radiation while being protected from damaging ultraviolet radiation. On Mars, the lack of an effective ozone shield allows ~30% more damaging ultraviolet radiation to reach the surface in comparison with Earth. However, our radiative transfer modelling shows that despite the intense surface ultraviolet radiation, there are radiatively habitable zones within exposed mid-latitude ice on Mars, at depths ranging from a few centimetres for ice with 0.01–0.1% dust, and up to a few metres within cleaner ice. Numerical models predict that dense dusty snow in the martian mid-latitudes can melt below the surface at present. Thus, if small amounts of liquid water are available at these depths, mid-latitude ice exposures could represent the most easily accessible locations to search for extant life on Mars.
Introduction
Dusty water ice (<~1% dust by mass) is exposed at the surface of Mars at latitudes poleward of 75° 1 , 2 . In the martian mid-latitudes ( \({{30}}{{-}}{{60}}{{^{\circ} \, }}\) latitude), dusty ice is typically overlain by desiccated material 3 , 4 , 5 . However, recent observations indicate that the overlying desiccated material can be removed by impact processes 6 or by slumping on steep slopes, to expose dusty ice at the surface in the mid-latitude regions of Mars 7 , 8 , 9 . The ice is thought to have been deposited as dusty snow during numerous periods of high obliquity that occurred over the last few million years 4 , 10 , 11 . While the precise grain size and density of exposed martian ice is currently unknown, snow on Mars is estimated to metamorphose into coarser-grained firn and eventually glacier ice on timescales ranging from decades/centuries 12 , 13 to millions of years 14 .
Snow, firn, and ice are very weakly absorbing at ultraviolet (0.2–0.4 \({{\mu }}{{\rm{m}}}\) ) and photosynthetically active radiation (PAR, 0.4–0.7 \({{\mu }}{{\rm{m}}}\) ) wavelengths, where solar radiation peaks, so the radiation is transmitted through these media to depths down to a few metres 15 , 16 , 17 , 18 , 19 , 20 . On Earth, organisms are protected from damaging ultraviolet radiation by atmospheric ozone and can harness the energy from PAR in ‘radiatively habitable zones’, even within ice 21 , 22 , 23 , 24 . But the lack of an effective ozone shield on Mars allows ~30% more damaging ultraviolet radiation to reach the surface of Mars than on Earth 25 .
Despite higher surface ultraviolet radiation levels, previous martian simulations suggested that radiatively habitable zones could exist within polar snow and ice on Mars, at depths between 0.05 and 4.5 m 26 , 27 , 28 . These radiatively habitable zones occur where damaging ultraviolet radiation is attenuated to safe levels while the flux of PAR is large enough for photosynthesis to occur. Previous martian studies used optical properties of various natural and synthetic snow analogues 26 , 28 , 29 and sea ice found on Earth 30 for their simulations. However, the optical properties used by these previous martian studies do not accurately represent the observed exposures of dusty ice on Mars, because martian snow has likely metamorphosed into coarser-grained firn or glacier ice 12 , 13 , 14 , which allow solar radiation to penetrate far deeper than snow 15 , 31 , and sea ice does not occur on Mars. Furthermore, polar locations on Mars are too cold for snow and ice to melt 12 , 32 . But exposed dusty ice in the mid-latitudes ( \({{30}}{{-}}{{50}}{{^{\circ} }}\) latitude) might be melting at present 4 , 9 , 12 , 32 .
To investigate whether radiatively habitable zones exist within mid-latitude ice exposures on Mars, we have developed a radiative transfer model for ice, based on work validated previously using measurements of snow, as well as snow that has metamorphosed into coarser-grained firn and glacier ice on Earth 33 , 34 , 35 , 36 , 37 and Mars 38 (see Methods). The model uses the Delta-Eddington method 39 , modified to account for varying refractive boundaries 40 , 41 , and can simulate vertically heterogeneous layers containing mixtures of snow, firn, ice, and impurities such as martian dust. We use the spectral solar radiation flux and incorporate the large (10 orders of magnitude) spectral variation in the imaginary (absorptive) part of the refractive index of H 2 O ice, which determines light absorption in the solar spectrum, in addition to accounting for absorption by martian dust. Martian dust is seven orders of magnitude more absorbing than ice (per unit mass) at visible and UV wavelengths, primarily due to the presence of ferric iron 42 , 43 ; it therefore significantly reduces both the ice’s albedo 38 and the penetration-depth of solar radiation 12 . Many previous models of martian ice 14 , 44 , 45 , 46 , 47 have ignored this spectral variation, either by assuming that all incoming solar radiation is absorbed within the topmost snow/ice model layer (typically a few mm) or by using a single broadband value, resulting in an underestimation of the amount of energy transmitted to depth up to a few orders of magnitude 15 . We therefore account for this large spectral variation, because it is crucial in accurately characterizing the transmission of radiation through snow and ice.
Spectral solar flux has not yet been measured within ice on Mars. Therefore, as an analogue for martian ice, we compare our model results to spectral irradiance measurements made within vertically heterogeneous glacier ice in Greenland containing impurities equivalent to small amounts ( ~ 1–2 parts per billion) of black carbon 17 . Overall, the model shows excellent agreement with the observations (Fig. 1 ), despite key input parameters such as grain size and impurity content not being measured directly.
The dashed curves represent results from our solar radiation model, and the solid-coloured curves represent measurements made in Greenland. Many existing ice models 14 , 44 , 45 , 46 , 47 either assume that no solar radiation penetrates below the top few millimeters, or use a single broadband extinction value indicated by the horizontal black line, which can lead to errors in the energy absorbed at depth up to a few orders of magnitude 15 .
Many snow and ice models for Mars 14 , 44 , 45 , 46 , 47 use a broadband approximation to calculate the fluxes of solar radiation at depth. As an illustrative example, Fig. 1 compares the Greenland measurements to the modelled broadband flux at 12 cm depth (black horizontal line) using an attenuation length of 0.8 m previously used by McKay 48 . The use of a constant attenuation length causes up to a ~ 60% underestimation of flux at wavelengths short of 0.7 \({{\rm{\mu }}}{{\rm{m}}}\) and up to a factor-of-17 overestimation at longer wavelengths where H 2 O ice is more absorptive.
Bohren and Barkstrom 49 showed that only 3% of the radiation scattered by a single ice grain is externally reflected, whereas 89% is transmitted after refraction and 8% is transmitted after internal reflections (their Eq. 24). So, as the grain radii and density of a snowpack increase over time, the number of scattering events at air-ice boundaries decreases. Thus, solar radiation penetrates to greater depth in coarse-grained snow, firn, and glacier ice than in fresh, fine-grained snow. Figure 2 shows the spectral variation of actinic flux with depth for dense snow, firn, and glacier ice on Mars. Actinic flux (sometimes referred to as scalar irradiance) is the radiance integrated over solid angle without cosine-weighting, summed over both upward and downward hemispheres and including both direct and diffuse radiation 50 . It is the quantity relevant for photosynthesis, because photosynthetic organisms utilize the photons available from all directions within snow and ice, independent of angle 28 , 51 . The results shown in Fig. 2 are for 33°S, the lowest latitude on Mars where H 2 O ice has been detected 9 .
Vertical panels are grouped by ice grain radius and density whereas horizontal panels are grouped by dust content (by mass). Note that the vertical (depth) scale is different in the three rows: 0–700 cm in the top row, 0–40 cm in the middle row, and 0–6 cm in the bottom row. The solar zenith angle is zero in all cases, and the peak solar flux over the course of the year is used.
In clean ice, the spectral distribution of subsurface actinic flux with depth mimics the spectral dependence of pure ice’s absorption coefficient, with the greatest penetration occurring at 0.42 \({{\mu }}{{\rm{m}}}\) , followed by wavelengths short of 0.4 \({{\mu }}{{\rm{m}}}\) . The greatest absorption within the PAR range occurs at wavelengths near 0.7 \({{\mu }}{{\rm{m}}}\) . In all cases, most of the solar radiation is attenuated within the top few metres for clean ice, with penetration increasing strongly with increasing grain size, because of the increased mean free path between scattering events. PAR penetrates deeply in clean ice, with radiation near 0.4 \({{\mu }}{{\rm{m}}}\) reaching down to ~6.5 metres (Fig. 2a–c and Supplementary Data Fig. 2a–c ). The weak absorption also leads to biologically damaging ultraviolet radiation (0.2–0.4 \({{\mu }}{{\rm{m}}}\) ) penetrating down to ~6 metres in clean ice.
Including small amounts of martian dust within the ice leads to two major changes in the modelled distribution of actinic flux at depth (Fig. 2d–i and Supplementary Data Fig. 2d–i ). The first change is to alter the spectral distribution of the actinic flux. The inclusion of even 0.01% dust causes the greatest penetration depth to shift toward long wavelengths near 0.7 \({{\mu }}{{\rm{m}}}\) , with the overall spectral distribution now following the spectral dependence of martian dust’s absorption coefficient. Between 0.2 and 0.4 \({{\mu }}{{\rm{m}}}\) martian dust is seven orders of magnitude more strongly absorbing than pure ice, which causes ultraviolet radiation to be attenuated more strongly by a factor of ~25. The second change is to drastically reduce the depth to which solar radiation can penetrate within the ice due to the combination of dust’s greater absorption and an increase in the number of scatterers within the ice. [Note the change of the depth scale in Fig. 2 (extending to 700 cm in a, b, c; 40 cm in d,e,f; 6 cm in g, h,i).] PAR penetrates an order of magnitude less with the inclusion of 0.01% dust, down to only 40 cm below the surface (Fig. 2f ). A further increase in dust concentration to 0.1% dust leads to a corresponding order of magnitude decrease in PAR penetration depth of down to just 5.5 cm below the surface (Fig. 2i ). The peak penetration depths for ultraviolet radiation are also reduced similarly to ~15 cm and 2 cm for ice containing 0.01% dust and 0.1% dust, respectively.
The penetration of actinic flux within martian ice leads to zones of radiative habitability that vary in depth and thickness by a few orders of magnitude depending on dust content, ice grain radius, latitude, and solar zenith angle, as shown in Fig. 3 and Supplementary Data Fig. 3 . Because there is some debate as to what the limit for PAR is on Earth, we present results using two commonly used PAR limits based on observations ( \({{10}}\,{{\rm{nmol\; photons}}} \, {{{\rm{m}}}}^{{{-}}{{2}}}\,{{{\rm{s}}}}^{{{-}}{{1}}}\) , equivalent to 2.2 m \({{\rm{W}}}\,{{{\rm{m}}}}^{{{-}}{{2}}}\) , and referred to here as the PAR lower limit) 52 and theoretical studies ( \({{20}}\,{{\mu }}{{\rm{mol\; photons}}} \, {{{\rm{m}}}}^{{{-}}{{2}}}\,{{{\rm{s}}}}^{{{-}}{{1}}}\) , equivalent to 4.353 \({{\rm{W}}}\,{{{\rm{m}}}}^{{{-}}{{2}}}\) , the PAR upper limit) 53 .
Zones where levels of DNA-damaging irradiance are safe and there is enough photosynthetically active radiation (PAR) for photosynthesis to occur are shown as the green areas, as a function of ( a ) dust content, ( b ) ice radius, ( c ) latitude, and ( d ) solar zenith angle. Two limits for PAR shown: a lower limit of \({{10}}\,{{\rm{nmol\; photons}}} \, {{{\rm{m}}}}^{{{-}}{{2}}}\,{{{\rm{s}}}}^{{{-}}{{1}}}\) (2.2 m \({{\rm{W}}}\,{{{\rm{m}}}}^{{{-}}{{2}}}\) ; red dashed lines), and an upper limit of \({{20}}\,{{\mu }}{{\rm{mol\; photons}}} \, {{{\rm{m}}}}^{{{-}}{{2}}}\,{{{\rm{s}}}}^{{{-}}{{1}}}\) (4.353 \({{\rm{W}}}\,{{{\rm{m}}}}^{{{-}}{{2}}}\) ; purple dotted lines). The DNA-damaging limit was taken to be \({{0}}{{.}}{{1}}\,{{\rm{W}}} \, {{{\rm{m}}}}^{{{-}}{{2}}}\) 25 , 26 . The reference values for the sensitivity analysis are ice grain radius 2.5 mm, density 775 \({{\rm{kg}}} \, {{{\rm{m}}}}^{{{-}}{{3}}}\) , 0.01% dust (by mass), solar zenith angle zero, and a latitude of 40°S. Note that the vertical (depth) axis in part ( a ) is on a log scale. In part ( c ), at each latitude the peak solar flux over the course of the year is used.
Figure 3a and Supplementary Data Fig. 3a show that for clean firn (10 −8 % dust, grain radius 2.5 mm, density 775 \({{\rm{kg}}} \, {{{\rm{m}}}}^{-3}\) ) at 40° latitude in the southern and northern hemispheres, respectively, a radiatively habitable zone is present between 2.15 and 3.10 metres using the lower PAR limit. However, no such radiatively habitable zone exists when the upper PAR limit is used. The effect of increasing the dust content of the ice is to move the zone of radiative habitability to shallower depths. For example, ice containing large amounts of dust (1%) has radiatively habitable zones between 2 and 6 mm for the lower PAR limit, or between 2 and 3 mm for the upper limit (Fig. 3a ). Lower values of dust content around 0.1% result in an order-of-magnitude increase in width of the radiatively habitable zone.
For ice containing 0.01% dust, an increase in grain radius leads to the radiatively habitable zone moving deeper and increasing in thickness (Fig. 3b and Supplementary Data Fig. 3b ). For coarse-grained snow ( ~ 360 \({{\mu }}{{\rm{m}}}\) , 390 \({{\rm{kg}}} \, {{{\rm{m}}}}^{-3}\) ), the radiatively habitable zone lies at 5 – 18 cm for the lower PAR limit and 5–7.5 cm for the upper limit. On the other extreme, for glacier ice with few bubbles (15 mm, 910 \({{\rm{kg}}} \, {{{\rm{m}}}}^{-3}\) ) the radiatively habitable zone depth ranges are 10–38 cm and 10–17 cm.
The effect of latitude on the radiatively habitable zones is relatively small because the thresholds for PAR and DNA damage are a few orders of magnitude smaller than the latitudinal variation of the peak annual solar fluxes reaching the martian surface. Thus, a change in latitude from 30° to 90° causes the depth at which the PAR upper limit is reached to shift by only a few centimetres (Fig. 3c and Supplementary Data Fig. 3c ).
The solar zenith angle, which varies diurnally and seasonally, also alters the solar fluxes at the surface. A larger solar zenith angle indicates the Sun is at a lower elevation. Increasing the solar zenith angle leads to a greater probability that an incoming photon will be scattered away from the surface, thereby reducing the depth to which solar radiation can penetrate below the ice surface. When the upper PAR limit is used, an increase in solar zenith angle from 0° to 60° causes the radiatively habitable zone to decrease in thickness from approximately 5 cm to 3 cm (Fig. 3d and Supplementary Data Fig. 3d ). By contrast, when the lower PAR limit is used, an increase in solar zenith angle has a relatively small impact on the radiatively habitable zones.
In conclusion, while the radiatively habitable zone of ice containing large amounts (1%) of dust is narrow (2 – 6 mm) and unlikely to support photosynthesis (Fig. 3a and Supplementary Data Fig. 3a ), its width increases by an order-of-magnitude at lower dust contents ( < 0.1%), where it overlaps with depth ranges imposed by other constraints (Fig. 3b–d and Supplementary Data Fig. 3b–d ).
Implications for life on Mars
Our analysis shows that despite higher surface ultraviolet radiation levels on Mars than on Earth, it is possible for terrestrial photosynthetic organisms to find locations within exposed ice on Mars with favourable solar radiative conditions. However, terrestrial microbes also require temperatures greater than 255 K for cell division to occur, and photosynthesis requires the presence of liquid water 54 . While martian polar ice deposits are too cold for melting to occur at these depths, and ice present within soil pores 55 is unlikely to melt 4 , 56 , numerical models developed specifically for martian snow 12 , 32 suggest that small amounts of melt and runoff (up to 0.33 mm per day for ~50 days each martian year) can be produced within exposed mid-latitude martian snowpacks containing ~0.1% dust, centimetres below the surface.
Although liquid water at the surface of Mars is highly likely to evaporate due to the low mean partial pressure of ~1 Pa for H 2 O vapor, liquid water produced within snow can be stable against evaporation. The gas within snow pores is typically saturated with water vapor 57 , and overlying snow/ice and dust can act as barriers against vapor diffusion to the dry martian atmosphere 4 , 12 , 45 , 57 , 58 , 59 , 60 . Thus, despite high energy losses from sublimation 61 , 62 , subsurface temperatures within dusty snow exposed on steep slopes can reach the melting point (273 K) 12 , 32 , where the local H 2 O partial pressure within snow is saturated at 611 Pa, allowing liquid water to be stable below the surface.
The predicted amounts of melt and runoff are very sensitive to the ice grain size and dust content 12 , 32 . The grain size of exposed ice in the martian mid-latitudes is currently unknown, but if the ice exposures on steep slopes 7 , 9 represent dusty firn or dusty glacier ice, terrestrial analogues suggest greater rates of subsurface melting than in snow because greater amounts of solar radiation penetrate below the surface (Fig. 2 ; refs. 31 , 63 ). The dust contents are also not well constrained for martian ice, but the ice is likely to contain less than 1% dust based on its albedo because ice with greater amounts of dust would be indistinguishable from nearby lithic material at wavelengths short of 1 \({{\mu }}{{\rm{m}}}\) 9 , 38 . However, as the ice melts and sublimates, it will gradually build up a lag of dust 64 , thereby increasing the dust content of the ice. Once the ice contains \({{\ge }}{{1}}{{ \% }}\) dust, the resulting opaque layer of dust can block incoming solar radiation and prevent subsurface melting 32 , 65 unless the dust lag layer forms below the surface 66 , in which case subsurface melting could be enhanced by increased solar absorption by a darker substrate within the ice 16 , 63 . Throughout the buildup of a surface dust lag, winds and surface frost-related processes can remove the overlying dust to re-expose the ice (as may be observed on steep slopes in the martian mid-latitudes 7 , 8 , 9 , 67 ) so that subsurface melting can occur once again under favourable conditions.
Under similar ephemeral near-freezing conditions, widespread microbial habitats are found in the shallow subsurface (top few centimetres to metres) of ice sheets, glaciers, and lake ice on Earth. Typically, these shallow subsurface habitats are formed by dark dust and sediment present at the surface of the ice absorbing greater amounts of solar radiation than the surrounding ice, heating up, and forming holes within the ice by melting 68 , 69 , 70 . These “cryoconite” holes continue to deepen until the amount of solar radiation transmitted to the dust and sediment is insufficient to melt further into the ice at some depth. Then, while liquid water is present around the sediment at the bottom of the hole, an ice lid forms at the surface, isolating the hole interior from the atmosphere. In the summer, photosynthesis occurs in the subsurface below the translucent ice lid, with nutrients scavenged from the dust and sediment present in the water-filled hole. During winter, the subsurface liquid refreezes, and photosynthesis ceases until the next summer.
A variety of organisms are found in these shallow subsurface habitats within terrestrial ice, such as cyanobacteria, chlorophytes, fungi, diatoms, and heterotrophic bacteria 71 . Typically, the most dominant organisms in these habitats are cyanobacteria 72 . Cyanobacteria have developed appropriate mechanisms to deal with a wide range of temperatures, nutrient deficiency, UV radiation, and desiccation 72 , 73 . These communities can persist even if temperatures are above freezing for only a few days per year 72 . The results presented here indicate that radiatively habitable zones exist within exposed mid-latitude ice on Mars at depths ranging from a few centimetres for ice with 0.01–0.1% dust, and up to a few metres within cleaner ice. Thus, if exposures of dusty mid-latitude martian ice are melting below the surface for a fraction of the year as predicted by numerical models 12 , 32 , then, like on Earth, microbes such as cyanobacteria could scavenge nutrients from the martian dust 74 mixed in with the ice, and utilize the small amounts of melt whilst being in a radiatively favourable habitat below the surface. These potential present-day habitats are centimetres to metres below the surface and could be the most easily accessible locations to find extant life on Mars via future robotic and human missions.
Downward solar spectrum
The Planetary Spectrum Generator 75 was used to simulate the downward spectral solar radiation flux at the surface of Mars. As inputs, the mean Mars-Sun distance of 1.52 AU was used with a solar zenith angle of 0 \({\scriptstyle{{\circ} \, \, \, }}\) , for a solar longitude of \({{0}}{{^{\circ} }}\) , at \({{0}}{{^{\circ} }} \, {{\rm{N}}}{{,}}\,{{0}}{{^{\circ} }} \, {{\rm{E}}}\) (using the ‘Looking up to the Sun’ viewing geometry). The Planetary Spectrum Generator’s standard martian atmospheric profile with no atmospheric aerosols (dust or water ice) was used. After dividing the downward solar spectrum ( \({{\rm{W}}} \, {{{\rm{m}}}}^{{{-}}{{2}}} \, {{\mu }}{{{\rm{m}}}}^{{{-}}{{1}}}\) ) by its spectrally integrated value ( \({{\rm{W}}} \, {{{\rm{m}}}}^{{{-}}{{2}}}\) ) we obtained the result shown in Supplementary Data Fig. 1 . Then, we multiplied peak annual broadband solar flux values ( \({{\rm{W}}} \, {{{\rm{m}}}}^{{{-}}{{2}}}\) ) from the validated Mars Climate Database 76 , 77 to obtain surface spectral fluxes for each latitude ( \({{\rm{W}}} \, {{{\rm{m}}}}^{{{-}}{{2}}}\,{{\mu }}{{{\rm{m}}}}^{{{-}}{{1}}}\) ).
Optical properties for dusty snow, firn and glacier ice on Mars
At present, there is some uncertainty about the precise H 2 O ice refractive index values at wavelengths short of 0.6 \({{\rm{\mu }}}{{\rm{m}}}\) . Thus, following Flanner, et al. 41 , we use a blend of values for the imaginary part of the complex refractive index of H 2 O ice. For 0.32 – 0.6 \({{\rm{\mu }}}{{\rm{m}}}\) we use values from Picard, et al. 78 ; for 0.6 – 3 \({{\rm{\mu }}}{{\rm{m}}}\) we use values from Warren and Brandt 79 . The Picard et al. 78 value at 0.32 \({{\rm{\mu }}}{{\rm{m}}}\) is used for 0.2 – 0.32 \({{\rm{\mu }}}{{\rm{m}}}\) . The use of this blend of imaginary refractive indices leads to relatively conservative estimates of solar radiation penetration depths when compared to results using the lower values of Warren and Brandt 79 . For the real part of the refractive index, we use values from Warren and Brandt 79 . For martian dust, we incorporate the spectral absorption properties derived by Wolff et al. 80 . Optical properties of snow, firn, and glacier ice mixtures with dust are computed using the method of Khuller et al. 38 .
Radiative transfer model for dusty snow, firn and glacier ice on Mars
Our radiative transfer model uses the solar transmission formulation of the Snow, Ice, and Aerosol Radiative Adding-Doubling model Version 4 SNICAR-ADv4; 35 . SNICAR-ADv4 is based on the Delta-Eddington method modified to include media with differing refractive indices such as snow, liquid water, or ice in the same vertical column 33 , 36 , 37 , 39 , 40 . The model outputs upward and downward spectral fluxes \({F}_{{\rm{up}}}\) and \({F}_{{\rm{down}}} \, ({{\rm{W}}} \, {{{\rm{m}}}}^{-2} \, {{\rm{\mu }}}{{{\rm{m}}}}^{-1})\) , which we convert into actinic flux. The total actinic flux within snow and ice is simply \(2({F}_{{\rm{up}}}+{F}_{{\rm{down}}})\) 50 , 51 .
Solar radiation results for Greenland glacier ice
The glacier ice in Greenland was measured at four depths for density, but neither the grain size nor the impurity content was measured 17 . Additionally, because these types of measurements within ice are challenging, other key uncertainties are present: (a) the density of the quasi-granular layer in the upper ~12 cm is somewhat uncertain, and (b) excess solar radiation could have entered from the side-wall of the experimental setup, into which the probe was inserted horizontally to make these measurements. Thus, we chose a plausible set of model inputs to match the measured values. Based on the measurements, the estimated ice absorption coefficient at \(0.4\,{{\rm{\mu }}}{{\rm{m}}}\) suggested that the ice contained the equivalent of ~1 ng/g of black carbon 17 , which is a typical value found on the Greenland Ice Sheet 63 . Thus, keeping this black-carbon concentration constant throughout, we used four layers in our model corresponding to each density measurement at depth, with the following grain radii, and densities for each layer: (1) 6.5 mm, 775 \({{\rm{kg}}} \, {{{\rm{m}}}}^{-3}\) , (2) 20 mm, 800 \({{\rm{kg}}} \, {{{\rm{m}}}}^{-3}\) , (3) 9 mm, 884 \({{\rm{kg}}} \, {{{\rm{m}}}}^{-3}\) , and (4) 7 mm, 888 \({{\rm{kg}}} \, {{{\rm{m}}}}^{-3}\) . The downward solar spectrum measured at the surface was used as input (grey upper curve in Fig. 1 ).
Photosynthetically Active Radiation (PAR)
Photosynthetically active radiation (PAR) is the spectrally integrated solar flux between 0.4 and 0.7 μm,
where both PAR ( \({{\rm{W}}}\,{{{\rm{m}}}}^{{{-}}{{2}}}\) ) and \({{S}}\left({{\lambda }}{{,}}{{z}}\right)\) , the spectral solar flux ( \({{\rm{W}}} \, {{{\rm{m}}}}^{{{-}}{{2}}}\,{{\mu }}{{{\rm{m}}}}^{{{-}}{{1}}}\) ), vary with depth \({{z}}\) (m). PAR is often expressed in units of \({{\rm{mol\; photons}}} \, {{{\rm{m}}}}^{{{-}}{{2}}}\,{{{\rm{s}}}}^{{{-}}{{1}}}\) , which can be converted to \({{\rm{W}}}\,{{{\rm{m}}}}^{{{-}}{{2}}}\) by a multiplier, \(\frac{{{hc}}{{{N}}}_{{{a}}}}{{{{\lambda }}}_{{{\rm{peak}}}}}\) , where \({{h}}\) is Planck’s constant, \({{c}}\) is the speed of light in vacuum, \({{{N}}}_{{{a}}}\) is Avogadro’s number, and \({{{\lambda }}}_{{{\rm{peak}}}}{{\approx }}{{0}}{{.}}{{55}}\,{{\mu }}{{\rm{m}}}\) is the wavelength of peak solar flux. We used values of 20 \({{\mu }}{{\rm{mol\; photons}}} \, {{{\rm{m}}}}^{{{-}}{{2}}}\,{{{\rm{s}}}}^{{{-}}{{1}}}\) (4.353 \({{\rm{W}}}\,{{{\rm{m}}}}^{{{-}}{{2}}}\) ) and 10 \({{\rm{nmol\; photons}}} \, {{{\rm{m}}}}^{{{-}}{{2}}}\,{{{\rm{s}}}}^{{{-}}{{1}}}\) (2.2 m \({{\rm{W}}}\,{{{\rm{m}}}}^{{{-}}{{2}}}\) ) as the upper and lower limits of PAR, based on observations and theory on Earth 26 , 52 , 53 .
DNA damaging irradiance
The DNA damaging irradiance at a given depth is given by the integral from 0.2 to 0.4 μm of the product of the spectral solar flux and the biological spectral weighting function for damage, \({{\varepsilon }}\left({{\lambda }}\right)\) :
The weighting function \({{\varepsilon }}\left({{\lambda }}\right)\) we use is taken from Cockell and Raven 26 . It was derived from a combination of DNA absorbance and the DNA action spectrum 81 , 82 . It has been used successfully to represent the relative amount of damage to DNA structures from ultraviolet radiation 26 , 27 , 28 . Following previous studies, we use a value of \({{0}}{{.}}{{1}}\,{{\rm{W}}} \, {{{\rm{m}}}}^{{{-}}{{2}}}\) as the irradiance limit for DNA damage, which is the terrestrial, equatorial value during vernal equinox, at midday 25 , 26 . This limit is conservative because numerous terrestrial organisms grow and survive under stronger DNA damage conditions 83 . Additionally, it may be possible for organisms to mutate and adapt to thrive under stronger DNA damage conditions, as is thought to have occurred on Archean Earth 83 .
Data availability
The spectral irradiance data 84 from the Greenland glacier ice study 17 is available at https://doi.org/10.1594/PANGAEA.930278 . Results from our radiative transfer modelling are available here 85 : https://doi.org/10.5281/zenodo.13277435 .
Code availability
The SNICAR-ADv4 code we use is available at https://github.com/chloewhicker/SNICAR-ADv4 .
Kieffer, H. H., Chase, S. C. Jr., Martin, T. Z., Miner, E. D. & Palluconi, F. D. Martian North Pole summer temperatures: Dirty water ice. Science 194 , 1341–1344 (1976).
Article CAS Google Scholar
Titus, T. N., Kieffer, H. H. & Christensen, P. R. Exposed water ice discovered near the south pole of Mars. Science 299 , 1048–1051 (2003).
Mustard, J. F., Christopher, D. C. & Moses, K. R. Evidence for recent climate change on Mars from the identification of youthful near-surface ground ice. Nature 412 , 411 (2001).
Christensen, P. Formation of recent martian gullies through melting of extensive water-rich snow deposits. Nature 422 , 45–48 (2003).
Plaut, J. J. et al. Radar evidence for ice in lobate debris aprons in the mid‐northern latitudes of Mars. Geophys. Res. Lett. 36 , https://doi.org/10.1029/2008GL036379 (2009).
Byrne, S. et al. Distribution of mid-latitude ground ice on Mars from new impact craters. Science 325 , 1674–1676 (2009).
Dundas, C. M. et al. Exposed subsurface ice sheets in the Martian mid-latitudes. Science 359 , 199 (2018).
Dundas, C. M. et al. The formation of gullies on Mars today. Geol. Soc., Lond., Spec. Publ. 467 , 67–94 (2019).
Article Google Scholar
Khuller, A. R. & Christensen, P. R. Evidence of Exposed Dusty Water Ice within Martian Gullies. J. Geophys. Res.: Planets , https://doi.org/10.1029/2020JE006539 (2021).
Jakosky, B. M. & Carr, M. A. Possible precipitation of ice at low latitudes of Mars during periods of high obliquity. Nature 315 , 559–561 (1985).
Madeleine, J. B. et al. Recent ice ages on Mars: the role of radiatively active clouds and cloud microphysics. Geophys. Res. Lett. 41 , 4873–4879 (2014).
Clow, G. D. Generation of liquid water on Mars through the melting of a dusty snowpack. Icarus 72 , 95–127 (1987).
Kieffer, H. H. H2O grain size and the amount of dust in Mars’ residual north polar cap. J. Geophys. Res.: Solid Earth 95 , 1481–1493 (1990).
Bramson, A., Byrne, S. & Bapst, J. Preservation of midlatitude ice sheets on Mars. J. Geophys. Res.: Planets 122 , 2250–2266 (2017).
Brandt, R. E. & Warren, S. G. Solar-heating rates and temperature profiles in Antarctic snow and ice. J. Glaciol. 39 , 99–110 (1993).
McKay, C., Clow, G., Andersen, D. & Wharton, R. Jr Light transmission and reflection in perennially ice‐covered Lake Hoare, Antarctica. J. Geophys. Res.: Oceans 99 , 20427–20444 (1994).
Cooper, M. G. et al. Spectral attenuation coefficients from measurements of light transmission in bare ice on the Greenland Ice Sheet. Cryosphere 15 , 1931–1953 (2021).
Schwerdtfeger, P. & Weller, G. The measurement of radiative and conductive heat transfer in ice and snow. Arch. f. ür. Meteorologie, Geophysik und Bioklimatologie, Ser. B 15 , 24–38 (1967).
Schwerdtfeger, P. Absorption, scattering and extinction of light in ice and snow. Nature 222 , 378–379 (1969).
Weller, G. E. Radiation diffusion in Antarctic ice media. Nature 221 , 355–356 (1969).
Rothschild, L. J. Earth analogs for Martian life. Microbes in evaporites, a new model system for life on Mars. Icarus 88 , 246–260 (1990).
Thomas, W. H., Duval, B. & Sierra, Nevada California, USA, snow algae: snow albedo changes, algal-bacterial interrelationships, and ultraviolet radiation effects. Arct. Alp. Res. 27 , 389–399 (1995).
Broady, P. A. Diversity, distribution and dispersal of Antarctic terrestrial algae. Biodivers. Conserv. 5 , 1307–1335 (1996).
Anesio, A. M., Lutz, S., Chrismas, N. A. & Benning, L. G. The microbiome of glaciers and ice sheets. npj Biofilms Microbiomes 3 , 10 (2017).
Córdoba-Jabonero, C., Lara, L., Mancho, A., Márquez, A. & Rodrigo, R. Solar ultraviolet transfer in the Martian atmosphere: biological and geological implications. Planet. Space Sci. 51 , 399–410 (2003).
Cockell, C. S. & Raven, J. A. Zones of photosynthetic potential on Mars and the early Earth. Icarus 169 , 300–310 (2004).
Córdoba-Jabonero, C., Zorzano, M.-P., Selsis, F., Patel, M. R. & Cockell, C. S. Radiative habitable zones in martian polar environments. Icarus 175 , 360–371 (2005).
France, J., King, M. & MacArthur, A. A photohabitable zone in the martian snowpack? A laboratory and radiative-transfer study of dusty water–ice snow. Icarus 207 , 133–139 (2010).
Perovich, D. K. Theoretical estimates of light reflection and transmission by spatially complex and temporally varying sea ice covers. J. Geophys. Res.: Oceans 95 , 9557–9567 (1990).
Grenfell, T. C. & Maykut, G. A. The optical properties of ice and snow in the Arctic Basin. J. Glaciol. 18 , 445–463 (1977).
Liston, G. E., Winther, J.-G., Bruland, O., Elvehøy, H. & Sand, K. Below-surface ice melt on the coastal Antarctic ice sheet. J. Glaciol. 45 , 273–285 (1999).
Williams, K. E., Toon, O. B., Heldmann, J. L., McKay, C. & Mellon, M. T. Stability of mid-latitude snowpacks on Mars. Icarus 196 , 565–577 (2008).
Mullen, P. C. & Warren, S. G. Theory of the optical properties of lake ice. J. Geophys. Res.: Atmospheres 93 , 8403–8414 (1988).
Dadic, R., Mullen, P. C., Schneebeli, M., Brandt, R. E. & Warren, S. G. Effects of bubbles, cracks, and volcanic tephra on the spectral albedo of bare ice near the Transantarctic Mountains: Implications for sea glaciers on Snowball Earth. J. Geophys. Res.: Earth Surf. 118 , 1658–1676 (2013).
Whicker, C. A. et al. SNICAR-ADv4: a physically based radiative transfer model to represent the spectral albedo of glacier ice. Cryosphere 16 , 1197–1220 (2022).
Wiscombe, W. J. & Warren, S. G. A model for the spectral albedo of snow.1. Pure snow. J. Atmos. Sci. 37 , 2712–2733 (1980).
Briegleb, P. & Light, B. A Delta-Eddington mutiple scattering parameterization for solar radiation in the sea ice component of the community climate system model. (National Center for Atmospheric Research, Boulder, Colorado, 2007).
Khuller, A. R., Christensen, P. R. & Warren, S. G. Spectral albedo of dusty martian H2O snow and ice. J. Geophys. Res.: Planets 126 , e2021JE006910 (2021).
Joseph, J. H., Wiscombe, W. & Weinman, J. The delta-Eddington approximation for radiative flux transfer. J. Atmos. Sci. 33 , 2452–2459 (1976).
Dang, C., Zender, C. S. & Flanner, M. G. Intercomparison and improvement of two-stream shortwave radiative transfer schemes in Earth system models for a unified treatment of cryospheric surfaces. Cryosphere 13 , 2325–2343 (2019).
Flanner, M. G. et al. SNICAR-ADv3: a community tool for modeling spectral snow albedo. Geoscientific Model Dev. 14 , 7673–7704 (2021).
Bell, J. F. et al. Mineralogic and compositional properties of martian soil and dust: Results from the Mars Pathfinder. J. Geophys. Res. 104 , 1721–1755 (2000).
Morris, R. V. et al. Mineralogy, composition, and alteration of Mars Pathfinder rocks and soils: Evidence from multispectural, elemental, and magnetic data on terrestrial analogue, SNC meteorite, and pathfinder samples. J. Geophys. Res. 105 , 1757–1817 (2000).
Kling, A. M., Haberle, R. M., McKay, C. P., Bristow, T. F. & Rivera-Hernandez, F. Subsistence of ice-covered lakes during the Hesperian at Gale crater. Mars. Icarus 338 , 113495 (2020).
Hecht, M. H. Metastability of liquid water on Mars. Icarus 156 , 373–386 (2002).
Dundas, C. M. & Byrne, S. Modeling sublimation of ice exposed by new impacts in the martian mid-latitudes. Icarus 206 , 716–728 (2010).
Schorghofer, N. Mars: Quantitative evaluation of crocus melting behind boulders. Astrophysical J. 890 , 49 (2020).
McKay, C. Thickness of tropical ice and photosynthesis on a snowball Earth. Geophys. Res. Lett. 27 , 2153–2156 (2000).
Bohren, C. F. & Barkstrom, B. R. Theory of the optical properties of snow. J. Geophys. Res. 79 , 4527–4535 (1974).
Madronich, S. Photodissociation in the atmosphere: 1. Actinic flux and the effects of ground reflections and clouds. J. Geophys. Res.: Atmospheres 92 , 9740–9752 (1987).
Simpson, W. R., King, M. D., Beine, H. J., Honrath, R. E. & Zhou, X. Radiation-transfer modeling of snow-pack photochemical processes during ALERT 2000. Atmos. Environ. 36 , 2663–2670 (2002).
Littler, M. M., Littler, D. S., Blair, S. M. & Norris, J. N. Deep-water plant communities from an uncharted seamount off San Salvador Island, Bahamas: distribution, abundance, and primary productivity. Deep Sea Res. Part A. Oceanographic Res. Pap. 33 , 881–892 (1986).
Raven, J. A., Kübler, J. E. & Beardall, J. Put out the light, and then put out the light. J. Mar. Biol. Assoc. U. Kingd. 80 , 1–25 (2000).
Rummel, J. D. et al. A new analysis of Mars “special regions”: findings of the second MEPAG Special Regions Science Analysis Group (SR-SAG2). Astrobiology 14 , 887–968 (2014).
Mellon, M. T. et al. Ground ice at the Phoenix landing site: Stability state and origin. J. Geophys. Res.: Planets 114 , https://doi.org/10.1029/2009JE003417 (2009).
Mellon, M. T., Sizemore, H. G., Heldmann, J. L., McKay, C. P. & Stoker, C. R. The habitability conditions of possible Mars landing sites for life exploration. Icarus 408 , 115836 (2024).
Neumann, T. A., Albert, M. R., Engel, C., Courville, Z. & Perron, F. Sublimation rate and the mass-transfer coefficient for snow sublimation. Int. J. Heat. Mass Transf. 52 , 309–315 (2009).
Farmer, C. B. Liquid water on Mars. Icarus 28 , 279–289 (1976).
Hudson, T. L. et al. Water vapor diffusion in Mars subsurface environments. J. Geophys. Res.: Planets 112 , https://doi.org/10.1029/2006JE002815 (2007).
Richardson, M. I. & Mischna, M. A. Long‐term evolution of transient liquid water on Mars. J. Geophys. Res.: Planets 110 , https://doi.org/10.1029/2004JE002367 (2005).
Ingersoll, A. P. Mars: Occurrence of liquid water. Science 168 , 972–973 (1970).
Khuller, A. R. & Clow, G. D. Turbulent Fluxes, Evaporation & Sublimation Rates on Earth, Mars, Titan & Exoplanets. J. Geophys. Res.: Planets , e2023JE008114, https://doi.org/10.1029/2023JE008114 (2024).
Paige, R. Sub-surface melt pools in the McMurdo Ice Shelf, Antarctica. J. Glaciol. 7 , 511–516 (1968).
Higuchi, K. & Nagoshi, A. Effect of particulate matter in surface snow layers on the albedo of perennial snow patches. IAHS AISH Publ. 118 , 95–97 (1977).
Google Scholar
Williams, K. E., Dundas, C. M. & Kahre, M. A. The formation mechanisms for mid-latitude ice scarps on Mars. Icarus 386 , 115174 (2022).
Doherty, S. J. et al. Observed vertical redistribution of black carbon and other insoluble light‐absorbing particles in melting snow. J. Geophys. Res.: Atmospheres 118 , 5553–5569 (2013).
Dundas, C. M. et al. Widespread Exposures of Extensive Clean Shallow Ice in the Mid‐Latitudes of Mars. J. Geophys. Res.: Planets , e2020JE006617, https://doi.org/10.1029/2020JE006617 (2021).
Wharton Jr, R. A., McKay, C. P., Simmons Jr, G. M. & Parker, B. C. Cryoconite holes on glaciers. Bioscience 35 , 499–503 (1985).
Fountain, A. G., Tranter, M., Nylen, T. H., Lewis, K. J. & Mueller, D. R. Evolution of cryoconite holes and their contribution to meltwater runoff from glaciers in the McMurdo Dry Valleys, Antarctica. J. Glaciol. 50 , 35–45 (2004).
Gordon, D., Priscu, J. & Giovannoni, S. Origin and phylogeny of microbes living in permanent Antarctic lake ice. Microb. Ecol. 39 , 197–202 (2000).
Cameron, K. A. et al. Supraglacial bacterial community structures vary across the Greenland ice sheet. FEMS Microbiol. Ecol. 92 , fiv164 (2016).
Vincent, W. F. in Algae and cyanobacteria in extreme environments (ed Joseph Seckbach) 287-301 (Springer, 2007).
Mueller, D. R., Vincent, W. F., Pollard, W. H. & Fritsen, C. H. Glacial cryoconite ecosystems: a bipolar comparison of algal communities and habitats. Nova Hedwig. Beih. 123 , 173–198 (2001).
Cockell, C. S. in Microbial life of the deep biosphere (eds Jens Kallmeyer & Dirk Wagner) 225-260 (De Gruyter, 2014).
Villanueva, G., Smith, M. D., Protopapa, S., Faggi, S. & Mandell, A. M. Planetary Spectrum Generator: An accurate online radiative transfer suite for atmospheres, comets, small bodies and exoplanets. J. Quant. Spectrosc. Radiat. Transf. 217 , 86–104 (2018).
Millour, E. et al. The Mars Climate Database (version 5.3). From Mars Express to ExoMars (2018).
Forget, F. et al. Improved general circulation models of the Martian atmosphere from the surface to above 80 km. J. Geophys. Res.: Planets 104 , 24155–24175 (1999).
Picard, G., Libois, Q. & Arnaud, L. Refinement of the ice absorption spectrum in the visible using radiance profile measurements in Antarctic snow. Cryosphere 10 , 2655–2672 (2016).
Warren, S. G. & Brandt, R. E. Optical constants of ice from the ultraviolet to the microwave: A revised compilation. J. Geophys. Res.: Atmospheres 113 , https://doi.org/10.1029/2007JD009744 (2008).
Wolff, M. et al. Wavelength dependence of dust aerosol single scattering albedo as observed by the Compact Reconnaissance Imaging Spectrometer. J. Geophys. Res.: Planets 114 , https://doi.org/10.1029/2009JE003350 (2009).
Horneck, G. Responses of Bacillus subtilis spores to space environment: results from experiments in space. Orig. Life Evolution Biosphere 23 , 37–52 (1993).
Green, A. & Miller, J. Measures of biologically effective radiation in the 280–340 nm region. CIAP Monogr. 5 , 2.60–70 (1975).
Cockell, C. S. Biological effects of high ultraviolet radiation on early Earth—a theoretical evaluation. J. Theor. Biol. 193 , 717–729 (1998).
Cooper, M. G. et al. Optical attenuation coefficients of glacier ice from 350–700 nm and raw irradiance values from 350–900 nm. (2021).
Khuller, A. R., Warren, S. G., Christensen, P. R. & Clow, G. D. (2024).
Download references
Acknowledgements
A part of this work was carried out at the Jet Propulsion Laboratory, California Institute of Technology, under a contract with the National Aeronautics and Space Administration. We would like to thank Joe Aslin, Candice Bedford, Kathleen Williamson, Matt Cooper, Carol Stoker, and one anonymous reviewer for very helpful feedback that greatly improved the paper. We would also like to thank Matt Cooper for providing the Greenland glacier radiation measurements, along with Erin Burkett, Scott Perl, Alejandro Martinez, Rahul Kushwaha, and Arnav Banerji for helpful discussions.
Author information
Authors and affiliations.
Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA, USA
Aditya R. Khuller
Department of Atmospheric Sciences, University of Washington, Seattle, WA, USA
Stephen G. Warren
School of Earth and Space Exploration, Arizona State University, Tempe, AZ, USA
Philip R. Christensen
Institute of Arctic and Alpine Research, University of Colorado Boulder, Boulder, CO, USA
Gary D. Clow
You can also search for this author in PubMed Google Scholar
Contributions
Radiative transfer calculations and analysis: A.R.K., S.G.W. Paper writing: A.R.K, S.G.W. Paper editing: A.R.K., S.G.W., P.R.C., G.D.C.
Corresponding author
Correspondence to Aditya R. Khuller .
Ethics declarations
Competing interests.
The authors declare no competing interests.
Peer review
Peer review information.
Communications Earth & Environment thanks Matthew Cooper, Carol Stoker and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Primary Handling Editors: Candice Bedford and Joe Aslin. A peer review file is available.
Additional information
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
Transparent peer review file, supplementary information, rights and permissions.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ .
Reprints and permissions
About this article
Cite this article.
Khuller, A.R., Warren, S.G., Christensen, P.R. et al. Potential for photosynthesis on Mars within snow and ice. Commun Earth Environ 5 , 583 (2024). https://doi.org/10.1038/s43247-024-01730-y
Download citation
Received : 23 February 2024
Accepted : 25 September 2024
Published : 17 October 2024
DOI : https://doi.org/10.1038/s43247-024-01730-y
Share this article
Anyone you share the following link with will be able to read this content:
Sorry, a shareable link is not currently available for this article.
Provided by the Springer Nature SharedIt content-sharing initiative
Quick links
- Explore articles by subject
- Guide to authors
- Editorial policies
Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.
share this!
October 17, 2024
This article has been reviewed according to Science X's editorial process and policies . Editors have highlighted the following attributes while ensuring the content's credibility:
fact-checked
peer-reviewed publication
trusted source
Could life exist below Mars ice? Study proposes possibilities
by Andrew Good, Karen Fox, Molly Wasser, NASA
While actual evidence for life on Mars has never been found, a new NASA study proposes microbes could find a potential home beneath frozen water on the planet's surface.
Through computer modeling, the study's authors have shown that the amount of sunlight that can shine through water ice would be enough for photosynthesis to occur in shallow pools of meltwater below the surface of that ice. Similar pools of water that form within ice on Earth have been found to teem with life, including algae, fungi, and microscopic cyanobacteria, all of which derive energy from photosynthesis.
"If we're trying to find life anywhere in the universe today, Martian ice exposures are probably one of the most accessible places we should be looking," said the paper's lead author, Aditya Khuller of NASA's Jet Propulsion Laboratory in Southern California.
Mars has two kinds of ice: frozen water and frozen carbon dioxide. For their paper, published in Communications Earth & Environment , Khuller and colleagues looked at water ice, large amounts of which formed from snow mixed with dust that fell on the surface during a series of Martian ice ages in the past million years. That ancient snow has since solidified into ice, still peppered with specks of dust.
Although dust particles may obscure light in deeper layers of the ice, they are key to explaining how subsurface pools of water could form within ice when exposed to the sun: Dark dust absorbs more sunlight than the surrounding ice, potentially causing the ice to warm up and melt up to a few feet below the surface.
Mars scientists are divided about whether ice can actually melt when exposed to the Martian surface. That's due to the planet's thin, dry atmosphere, where water ice is believed to sublimate—turn directly into gas—the way dry ice does on Earth. But the atmospheric effects that make melting difficult on the Martian surface wouldn't apply below the surface of a dusty snowpack or glacier.
Thriving microcosms
On Earth, dust within ice can create what are called cryoconite holes—small cavities that form in ice when particles of windblown dust (called cryoconite) land there, absorb sunlight, and melt farther into the ice each summer. Eventually, as these dust particles travel farther from the sun's rays, they stop sinking, but they still generate enough warmth to create a pocket of meltwater around them. The pockets can nourish a thriving ecosystem for simple lifeforms..
"This is a common phenomenon on Earth," said co-author Phil Christensen of Arizona State University in Tempe, referring to ice melting from within. "Dense snow and ice can melt from the inside out, letting in sunlight that warms it like a greenhouse, rather than melting from the top down."
Christensen has studied ice on Mars for decades. He leads operations for a heat-sensitive camera called THEMIS (Thermal Emission Imaging System) aboard NASA's 2001 Mars Odyssey orbiter. In past research, Christensen and Gary Clow of the University of Colorado Boulder used modeling to demonstrate how liquid water could form within dusty snowpack on the Red Planet. That work, in turn, provided a foundation for the new paper focused on whether photosynthesis could be possible on Mars.
In 2021, Christensen and Khuller co-authored a paper on the discovery of dusty water ice exposed within gullies on Mars, proposing that many Martian gullies form by erosion caused by the ice melting to form liquid water.
This new paper suggests that dusty ice lets in enough light for photosynthesis to occur as deep as 9 feet (3 meters) below the surface . In this scenario, the upper layers of ice prevent the shallow subsurface pools of water from evaporating while also providing protection from harmful radiation. That's important, because unlike Earth, Mars lacks a protective magnetic field to shield it from both the sun and radioactive cosmic ray particles zipping around space.
The study authors say the water ice that would be most likely to form subsurface pools would exist in Mars' tropics, between 30 degrees and 60 degrees latitude, in both the northern and southern hemispheres.
Khuller next hopes to re-create some of Mars' dusty ice in a lab to study it up close. Meanwhile, he and other scientists are beginning to map out the most likely spots on Mars to look for shallow meltwater—locations that could be scientific targets for possible human and robotic missions in the future.
Journal information: Nature Communications Earth & Environment , Communications Earth & Environment
Provided by NASA
Explore further
Feedback to editors
Study: Smaller, more specific academic journals hold more sway over conservation policy
3 hours ago
A near-Earth microquasar emerges as a source of powerful radiation
Imaging X-ray Polarimetry Explorer helps researchers determine shape of black hole corona
4 hours ago
Environmental DNA and epidemics in wood frogs: Collaboration examines eDNA's precision in population size estimation
More efficient phenotypic screening method can simultaneously test multiple drugs
Pioneering robot system enables 24/7 monitoring of honeybee behavior
DNA-binding C2H2 zinc finger proteins also regulate RNA processing, researchers discover
Harnessing diamond imperfections opens a new frontier in quantum sensor development
Catching prey with grappling hooks and cannons: The unusual weapons arsenal of a predatory marine bacterium
Philosopher finds glitch in worldwide patent laws
Relevant physicsforums posts, our beautiful universe - photos and videos.
7 hours ago
How feasible is home radio astronomy?
Oct 16, 2024
Solar Activity and Space Weather Update thread
Oct 11, 2024
Exploring the Sun: Amateur Solar Imaging Techniques
Oct 10, 2024
Optimizing Exposure Times: Balancing Efficiency and Image Quality
Oct 8, 2024
Strange cosmic particles in my detector
Oct 7, 2024
More from Astronomy and Astrophysics
Related Stories
Martian snow is dusty, could potentially melt, new study shows
Aug 23, 2021
Melting dusty ice may have carved Martian gullies
Feb 15, 2021
Gullies on Mars could have been formed by recent periods of liquid meltwater, study suggests
Jun 29, 2023
Image: Ice flows on Mars
Dec 19, 2023
Study looks more closely at Mars' underground water signals
Jun 25, 2021
Citizen scientists help map ridge networks on Mars
Apr 6, 2022
Recommended for you
Second exoplanet detected orbiting an early G-type star
9 hours ago
Unlocking cosmic origins: Researchers trace 70% of meteorites to 3 asteroid families
It's twins! Mystery of famed brown dwarf solved
Audible storm waves could turbocharge Earth's radiation belts
Scientists date moon's oldest impact basin to over 4.32 billion years ago
October's supermoon pairs with a comet for a special nighttime spectacle
Let us know if there is a problem with our content.
Use this form if you have come across a typo, inaccuracy or would like to send an edit request for the content on this page. For general inquiries, please use our contact form . For general feedback, use the public comments section below (please adhere to guidelines ).
Please select the most appropriate category to facilitate processing of your request
Thank you for taking time to provide your feedback to the editors.
Your feedback is important to us. However, we do not guarantee individual replies due to the high volume of messages.
E-mail the story
Your email address is used only to let the recipient know who sent the email. Neither your address nor the recipient's address will be used for any other purpose. The information you enter will appear in your e-mail message and is not retained by Phys.org in any form.
Newsletter sign up
Get weekly and/or daily updates delivered to your inbox. You can unsubscribe at any time and we'll never share your details to third parties.
More information Privacy policy
Donate and enjoy an ad-free experience
We keep our content available to everyone. Consider supporting Science X's mission by getting a premium account.
E-mail newsletter
An official website of the United States government
Official websites use .gov A .gov website belongs to an official government organization in the United States.
Secure .gov websites use HTTPS A lock ( Lock Locked padlock icon ) or https:// means you've safely connected to the .gov website. Share sensitive information only on official, secure websites.
- Publications
- Account settings
- Advanced Search
- Journal List
Photosynthetic Research in Plant Science
Ayumi tanaka, amane makino.
- Author information
- Article notes
- Copyright and License information
Issue date 2009 Apr.
The online version of this article has been published under an open access model. Users are entitled to use, reproduce, disseminate, or display the open access version of this article for non-commercial purposes provided that: the original authorship is properly and fully attributed; the Journal and the Japanese Society of Plant Physiologists are attributed as the original place of publication with the correct citation details given; if an article is subsequently reproduced or disseminated not in its entirety but only in part or as a derivative work this must be clearly indicated. For commercial re-use, please contact [email protected]
Photosynthesis is a highly regulated, multistep process. It encompasses the harvest of solar energy, transfer of excitation energy, energy conversion, electron transfer from water to NADP + , ATP generation and a series of enzymatic reactions that assimilate carbon dioxide and synthesize carbohydrate.
Photosynthesis has a unique place in the history of plant science, as its central concepts were established by the middle of the last century, and the detailed mechanisms have since been elucidated. For example, measurements of photosynthetic efficiency (quantum yield) at different wavelengths of light (Emerson and Lews 1943 ) led to the insight that two distinct forms of Chl must be excited in oxygenic photosynthesis. These results suggested the concept of two photochemical systems. The reaction center pigments of PSII and PSI (P680 and P700, respectively) were found by studying changes in light absorbance in the red region (Kok 1959 , Döring et al. 1969 ). Chls with absorbance maxima corresponding to these specific wavelengths were proposed as the final light sink. These Chls were shown to drive electron transfer by charge separation. The linkage of electron transfer and CO 2 assimilation was suggested by studies on Hill oxidant (Hill 1937 ). A linear electron transport system with two light-driven reactions (Z scheme) was proposed based upon observations of the redox state of cytochromes (Hill and Bendall 1960 , Duysens et al. 1961 ), and photophosphorylation was found to be associated with thylakoid fragments (Arnon et al. 1954 ). The metabolic pathway that assimilates carbon by fixation of CO 2 was discovered by Calvin's group who used 14 CO 2 radioactive tracers in the 1950s (Bassham and Calvin 1957 ). This was the first significant discovery in biochemistry made using radioactive tracers. The primary reaction of CO 2 fixation is catalyzed by Rubisco (Weissbach et al. 1956 ), initially called Fraction 1 protein (Wildman and Bonner 1947 ). Rubisco is the most abundant protein in the world, largely because it is also the most inefficient with the lowest catalytic turnover rate (1–3 s –1 ). Another CO 2 fixation pathway was then found in sugarcane (Kortschak et al. 1964, Hatch and Slack 1965) and named C 4 photosynthesis.
Although photosynthesis plays the central role in the energy metabolism of plants, historically there have not been strong interactions between photosynthesis research and other fields of plant science. Many techniques and tools developed for photosynthesis research have not been widely used in other fields because they were developed to examine phenomena unique to photosynthesis. For example, excitation energy transfer and charge separation are fundamental but unique processes of photosynthesis. Another reason for the historic isolation of photosynthesis research within plant science is that it was long believed that CO 2 fixation and carbohydrate production are the sole function of photosynthesis, with carbohydrates representing the only link between photosynthesis and other biological phenomena.
However, this situation has begun to change. Recent research has revealed that photosynthesis is closely related to a variety of other physiological processes. It is a major system for controlling the redox state of cells, playing an important role in regulating enzyme activity and many other cellular processes (Buchanan and Balmer 2005 , Hisabori et al. 2007 ). Photosynthesis also generates reactive oxygen species, which are now appreciated as being regulatory factors for many biological processes rather than inevitable by-products of photosynthesis (Wagner et al. 2004 , Beck 2005 ). Precursor molecules of Chl, which are a major component of photosynthesis, act as a chloroplast-derived signal, and are involved in regulating the cell cycle (Kobayashi et al. 2009 ). In light of this new information, it seems important to re-evaluate the function(s), both potential and demonstrated, of photosynthesis from a variety of view points. Photosynthesis research now employs the methods and tools of molecular biology and genetics, which are central methods for plant science in general. Meanwhile, Chl fluorescence and gas exchange measurements, developed especially for photosynthesis research, are now widely used in stress biology and ecology.
Photosynthesis research also contributes to our understanding of ecological phenomena and even the global environments (Farquhar et al. 1980 , de Pury and Farquhar 1997 , Monsi and Saeki 2005 ). Indeed, photosynthesis is now an integral component of simulation models used to predict the future of our planet. Improving the efficiency of photosynthesis by artificial modification of photosynthetic proteins and pathways has long been considered impossible or at best problematic, because, over evolutionary time, photosynthesis has become complex and tightly regulated. However, recent advances have made it possible to manipulate photosynthesis using molecular genetic technology (Andrews and Whiney 2003 , Raines 2006 ). These advances may have positive influences on crop productivity (Parry et al. 2007 ) as photosynthetic rates have frequently been correlated with biomass accretion (Kruger and Volin 2006 ). Thus, we can expect many more novel concepts to be added to this history of photosynthetic research.
As photosynthesis research tackles new challenges, we should also continue to re-evaluate past research. Oxygen evolution, energy dissipation and cyclic electron transport are crucial processes during photosynthesis, yet their mechanisms still remain to be clarified. We have very limited knowledge of the formation and degradation of photosynthetic apparatus. Also, although photosynthesis plays a central role in C and N metabolism in plants, we do not yet understand how potential photosynthesis is related to crop productivity.
Plant and Cell Physiology would like to contribute to the development of novel concepts, pioneering new fields and solving the unanswered questions of photosynthesis. This special issue covers a wide range of topics in photosynthesis research. Terashima et al. (pp. 684–697) readdress the enigmatic question of why leaves are green. They show that the light profile through a leaf is steeper than that of photosynthesis, and that the green wavelengths in white light are more effective in driving photosynthesis than red light. Evans (pp. 698–706) proposes a new model using Chl fluorescence to explore modifications in quantum yield with leaf depth. This new multilayered model can be applied to study variations in light absorption profiles, photosynthetic capacity and calculation of chloroplastic CO 2 concentration at different depths through the leaf.
Singlet oxygen, 1 O 2 , is produced by the photosystem and Chl pigments. 1 O 2 not only causes physiological damage but also activates stress response programs. The flu mutant of Arabidopsis thaliana overaccumulates protochlorophyllide that upon illumination generates singlet oxygen, causing growth cessation and cell death. Coll et al. (pp. 707–718) have isolated suppressor mutants, dubbed ‘singlet oxygen-linked death activator’ (soldat), that specifically abrogate 1 O 2 -mediated stress responses in young flu seedlings, and they discuss the processes of acclimation to stresses. Phephorbide a is a degradation product of Chl and one of the most powerful photosensitzing molecules. Mutants defective in pheophorbide a oxygenase, which converts phephorbide a to open tetrapyrrole, accumulate pheophorbide a and display cell death in a light-dependent manner. Hirashima et al. (pp. 719–729) report that pheophorbide a is involved in this light-independent cell death.
Plants regulate the redox level of the plastoquinone pool in response to the light environment. In acclimation to high-light conditions, the redox level is kept in an oxidized state by the plastoquinone oxidation system (POS). Miyake et al. (pp. 730–743) investigated the mechanism of POS using the Chl fluorescence parameter, qL.
Nagai and Makino (pp. 744–755) examine in detail the differences between rice and wheat, the two most commercially important crops, in the temperature responses of CO 2 assimilation and plant growth. They find that the difference in biomass production between the two species at the level of the whole plant depends on the difference in N-use efficiency in leaf photosynthesis and growth rate. Sage and Sage (pp. 756–772) examine chlorenchyma structure in rice and related Oryza species in relation to photosynthetic function. They find that rice chlorenchyma architecture includes adaptations to maximize the scavenging of photorespired CO 2 and to enhance the diffusive conductance of CO 2 . In addition, they consider that the introduction of Kranz anatomy does not require radical anatomical alterations in engineering C 4 rice.
Bioinformatics has become a powerful tool, especially in photosynthetic research, because photosynthetic organisms have a wide taxonomic distribution among prokaryotes and eukaryotes. Ishikawa et al. (pp. 773–788) present the results of a pilot study of functional orthogenomics, combining bioinformatic and experimental analyses to identify nuclear-encoded chloroplast proteins of endosymbiontic origin (CRENDOs). They conclude that phylogenetic profiling is useful in finding CPRENDOs, although the physiological functions of orthologous genes may be different in chloroplasts and cyanobacteria.
We hope you enjoy this special issue, and would like to invite you to submit more excellent papers to Plant and Cell Physiology in the field of photosynthesis.
- Andrews TJ, Whitney SM. Manipulating ribulose bisphosphate carboxylase/oxygenase in the chloroplasts of higher plants. Arch. Biochem. Biophys. 2003;414:159–169. doi: 10.1016/s0003-9861(03)00100-0. [ DOI ] [ PubMed ] [ Google Scholar ]
- Arnon DI, Allen MB, Whatley FR. Photosynthesis by isolated chloroplasts. Nature. 1954;174:394–396. doi: 10.1038/174394a0. [ DOI ] [ PubMed ] [ Google Scholar ]
- Bassham JA, Calvin M. The Path of Carbon in Photosynthesis. New Jersey: Prentice-Hall; 1957. [ Google Scholar ]
- Beck CF. Signaling pathways from the chloroplast to the nucleus. Planta. 2005;222:743–756. doi: 10.1007/s00425-005-0021-2. [ DOI ] [ PubMed ] [ Google Scholar ]
- Buchanan BB, Balmer Y. Redox regulation: a broadening horizon. Annu. Rev. Plant Biol. 2005;56:187–220. doi: 10.1146/annurev.arplant.56.032604.144246. [ DOI ] [ PubMed ] [ Google Scholar ]
- Coll NS, Danon A, Meurer J, Cho WK, Apel K. Characterization of soldat8, a suppressor of singlet oxygen-induced cell death in Arabidopsis seedlings. Plant Cell Physiol. 2009;50:707–718. doi: 10.1093/pcp/pcp036. [ DOI ] [ PubMed ] [ Google Scholar ]
- de Pury DGG, Farquhar GD. Simple scalling of photo-synthesis from leaves to canopies without the errors of big-leaf models. Plant Cell Environ. 1997;20:537–557. [ Google Scholar ]
- Döring G, Renger G, Vater J, Witt HT. Properties of the photoactive chlorophyll-aII in photosynthesis. Z. Naturforsch. B. 1969;24:1139–1143. doi: 10.1515/znb-1969-0911. [ DOI ] [ PubMed ] [ Google Scholar ]
- Duysens JNM, Amesz J, Kamp BM. Two photochemical systems in photosynthesis. Nature. 1961;190:510–511. doi: 10.1038/190510a0. [ DOI ] [ PubMed ] [ Google Scholar ]
- Emerson R, Lews ML. The dependence of the quantum yield of chlorella photosynthesis on wave length of light. Amer. J. Bot. 1943;30:165–178. [ Google Scholar ]
- Evans JR. Exploring chlorophyll fluorescence with a multilayer leaf model. Plant Cell Physiol. 2009;50:698–706. doi: 10.1093/pcp/pcp041. [ DOI ] [ PubMed ] [ Google Scholar ]
- Farquhar GD, von Caemmerer S, Berry JA. A biochemical model of photosynthetic CO2 assimilation in leaves of C3 species. Planta. 1980;149:78–90. doi: 10.1007/BF00386231. [ DOI ] [ PubMed ] [ Google Scholar ]
- Hatch MD, Slack CR. Photosynthesis by sugarcane leaves. A new carboxylation reaction and the pathway of sugar formation. Biochem. J. 1966;101:103–111. doi: 10.1042/bj1010103. [ DOI ] [ PMC free article ] [ PubMed ] [ Google Scholar ]
- Hill R. Oxygen evolved by isolated chloroplasts. Nature. 1937;139:881–882. [ Google Scholar ]
- Hill R, Bendall F. Function of the two cytochrome components in chloroplasts: a working hypothesis. Nature. 1960;186:136–137. [ Google Scholar ]
- Hirashima M, Tanaka R, Tanaka A. Light-independent cell death induced by accumulation of pheophorbide a in Arabidopsis thaliana. Plant Cell Physiol. 2009;50:719–729. doi: 10.1093/pcp/pcp035. [ DOI ] [ PubMed ] [ Google Scholar ]
- Hisabori T, Motohashi K, Hosoya-Matsuda N, Ueoka-Nakanishi H, Romano PG. Towards a functional dissection of thioredoxin networks in plant cells. Photochem. Photobiol. 2007;83:145–151. doi: 10.1562/2006-02-27-IR-816. [ DOI ] [ PubMed ] [ Google Scholar ]
- Ishikawa M, Fujiwara M, Sonoike K, Sato N. Orthogeno-mics of photosynthetic organisms: bioinformatic and experimental analysis of chloroplast proteins of endosymbiotic origin in Arabidopsis and their counterparts in Synechocystis. Plant Cell Physiol. 2009;50:773–788. doi: 10.1093/pcp/pcp027. [ DOI ] [ PubMed ] [ Google Scholar ]
- Kobayashi Y, Kanesaki Y, Tanaka A, Kuroiwa H, Kuroiwa T, Tanaka K. Tetrapyrrole signal as a cell-cycle coordinator from organelle to nuclear DNA replication in plant cells. Proc. Natl Acad. Sci. USA. 2009;106:803–807. doi: 10.1073/pnas.0804270105. [ DOI ] [ PMC free article ] [ PubMed ] [ Google Scholar ]
- Kok B. Light induced absorption changes in photosynthetic organisms. II. A split-beam difference spectrophotometer. Plant Physiol. 1959;34:184–192. doi: 10.1104/pp.34.3.184. [ DOI ] [ PMC free article ] [ PubMed ] [ Google Scholar ]
- Kortschak HP, Harrt CE, Burr GO. Carbon dioxide fixation in sugarcane leaves. Plant Physiol. 1965;40:209–213. doi: 10.1104/pp.40.2.209. [ DOI ] [ PMC free article ] [ PubMed ] [ Google Scholar ]
- Kruger EL, Volin JC. Reexamining the empirical relation between plant growth and leaf photosynthesis. Funct. Plant Biol. 2006;33:421–429. doi: 10.1071/FP05310. [ DOI ] [ PubMed ] [ Google Scholar ]
- Miyake C, Amako K, Shiraishi N, Sugimoto T. Acclimation of tobacco leaves to high light intensity drives the plastoquinone oxidation system—relationship among the fraction of open PSII centers, non-photochemical quenching of Chl fluorescence and the maximum quantum yield of PSII in the dark. Plant Cell Physiol. 2009;50:730–743. doi: 10.1093/pcp/pcp032. [ DOI ] [ PubMed ] [ Google Scholar ]
- Monsi M, Saeki T. On the factor light in plant communities and its importance for matter production. (English version of Jpn. J. Bot. 14: 22–52, 1953). Ann. Bot. 2005;95:549–567. doi: 10.1093/aob/mci052. [ DOI ] [ PMC free article ] [ PubMed ] [ Google Scholar ]
- Nagai T, Makino A. Differences between rice and wheat in temperature responses of photosynthesis and plant growth. Plant Cell Physiol. 2009;50:744–755. doi: 10.1093/pcp/pcp029. [ DOI ] [ PMC free article ] [ PubMed ] [ Google Scholar ]
- Parry MAJ, Madgwick PJ, Carvalho JFC, Andralojc PJ. Prospects from increasing photosynthesis by overcoming the limitations of Rubisco. J. Agric. Sci. 2007;145:31–43. [ Google Scholar ]
- Raines CA. Transgenic approaches to manipulate the environmental responses of the C-3 carbon fixation cycle. Plant Cell Environ. 2006;29:331–339. doi: 10.1111/j.1365-3040.2005.01488.x. [ DOI ] [ PubMed ] [ Google Scholar ]
- Sage TL, Sage RF. The functional anatomy of rice leaves: implications for refixation of photorespiratory CO2 and efforts to engineer C4 photosynthesis into rice. Plant Cell Physiol. 2009;50:756–772. doi: 10.1093/pcp/pcp033. [ DOI ] [ PubMed ] [ Google Scholar ]
- Terashima I, Fujita T, Inoue T, Chow WS, Oguchi R. Green light drives leaf photosynthesis more efficiently than red light in strong white light: revisiting the enigmatic question of why leaves are green. Plant Cell Physiol. 2009;50:684–697. doi: 10.1093/pcp/pcp034. [ DOI ] [ PubMed ] [ Google Scholar ]
- Wagner D, Przybyla D, Op den Camp R, Kim C, Landgraf F, Lee KP, et al. The genetic basis of singlet oxygen-induced stress responses of Arabidopsis thaliana. Science. 2004;306:1183–1185. doi: 10.1126/science.1103178. [ DOI ] [ PubMed ] [ Google Scholar ]
- Weissbach A, Horecker BL, Hurwitz J. The enzymatic formation of phosphoglyceric acid from ribulose diphosphate and carbon dioxide. J. Biol. Chem. 1956;218:795–810. [ PubMed ] [ Google Scholar ]
- Wildman SG, Bonner J. The proteins of green leaves. I. Isolation, enzymic properties, and auxin contents of spinach cytoplasmic proteins. Arch. Biochem. 1947;14:381–413. [ PubMed ] [ Google Scholar ]
- View on publisher site
- PDF (127.9 KB)
- Collections
Similar articles
Cited by other articles, links to ncbi databases.
- Download .nbib .nbib
- Format: AMA APA MLA NLM
IMAGES
VIDEO
COMMENTS
In this paper, we present a comprehensive study to explore potential interactive effect of light intensity and light quality on C 3 photosynthesis and underlying processes. We quantified the photosynthetic response of plants to blue, green, and red light over a wide PPFD range to better describe how light intensity and waveband interact.
These studies have become the foundation for our plant lighting research as light emitting diodes ... Chow W. S., and Oguchi R., "Green light drives leaf photosynthesis more efficiently than red light in strong white light: revisiting the enigmatic question of why leaves are green," Plant and cell physiology, vol. 50, no. 4, pp. 684-697, ...
(a) Light-emitting diode light(s) can sustain normal plant growth. Pioneer experiments on plant growth under red LEDs on lettuce were reported by Bula et al.[]Martineau et al. [] calculated that the amounts of dry matter per mole of artificial lighting gained by lettuce grown using red (650 nm) LEDs or high-pressure sodium lamps were identical, and Chang et al.
In light of the (expected) impact of thylakoid electron transport rate, the rate of regulation of the light reactions and the longevity of the photosynthetic activity over the growing season, I will focus in this review on approaches that aim to modify the light reactions of photosynthesis, i.e., on attempts to increase ε conversion (for an overview see Figure 1) and only refer in passing to ...
1. Introduction. Photosynthesis is a complex metabolic process, in which the energy of light, captured by an elaborate system of pigment-containing proteins, is used for the reduction of CO 2, which is an essential step in the biosynthesis of organic compounds.As photosynthesis uses the most abundant energy source available on Earth, it is the most important biochemical process in biological ...
We attribute this feature to the ground-state bleaching of the photoactive reaction centres of PSI and PSII. A lower-energy feature at 715 nm grew within roughly 2 ps after photo-excitation and ...
The impacts of wavelengths in 500-600 nm on plant response and their underlying mechanisms remain elusive and required further investigation. Here, we investigated the effect of light quality on leaf area growth, biomass, pigments content, and net photosynthetic rate (Pn) across three Arabidopsis thaliana accessions, along with changes in transcription, photosynthates content, and ...
Photosynthesis Research is an international journal publishing research in all areas related to photosynthesis. Discusses both basic and applied aspects of photosynthesis. Welcomes research in all photosynthetic systems, including natural organisms and biomimetic systems. Covers a broad range of topics including photophosphorylation, carbon ...
Photosynthesis is a complex metabolic process in which solar energy is utilized to convert atmospheric carbon dioxide (CO 2) into organic compounds.Traditionally, it is divided into two phases: (1) the light energy captured by the pigment-protein complexes is exchanged into energy-rich bonds of molecules such as ATP and NADPH, and (2) when ATP and NADPH drive the fixation of CO 2 in ...
The process of. photosynthesis in plants is based on two reactions that are carried out by separate parts. of the chloroplast. The light reactions occur in the chloroplast thylakoid membrane and ...
The idea of two light reactions and two types of PSs had its beginning in the 1943 experiments of Robert Emerson and Charleton Lewis on the 'red drop' in the action spectrum of the quantum yield of photosynthesis (Emerson and Lewis, 1943) and in the 1957 'Emerson enhancement' effect, that is when the rate of photosynthesis in two lights ...
The impacts of wavelengths in 500-600 nm on plant response and their underlying mechanisms remain elusive and required further investigation. Here, we investigated the effect of light quality on leaf area growth, biomass, pigments content, and net photosynthetic rate (Pn) across three Arabidopsis thaliana accessions, along with changes in transcription, photosynthates content, and ...
Photosynthesis is one of the most incomparable and meticulous metabolic processes that maximize the use of. available light, carbon and nitrogen and minimizes th e destructive effects of surplus ...
Abstract. Photosynthesis is defined as a light-dependent process; however, it is negatively influenced by high light (HL) intensities. To investigate whether the memory of growth under monochromatic or combinational lights can influence plant responses to HL, rose plants were grown under different light spectra [including red (R), blue (B), 70:30 % red:blue (RB) and white (W)] and were exposed ...
Photosynthesis is the process by which photosynthetic organisms utilize the energy from sunlight to assimilate CO 2 from the atmosphere and convert it into soluble carbohydrates, which are then used for plant growth (Bassham and Calvin 1960; Biel and Fomina 2015; Calvin and Benson 1948; Raines 2003).To meet global demand, the predicted requirement to increase crop yield, including the global ...
PAR intensity is an important factor that determines the rate of photosynthesis. Too high or too low PAR intensities adversely affect the photosynthetic machinery. At low light intensities above the light compensation point (LCP), photosynthetic rate increases proportionally to the light intensity and reaches a maximum.
Introduction. The photosynthetic activity of light is wavelength dependent. Based on McCree's work (McCree, 1971, 1972), photosynthetically active radiation is typically defined as light with a wavelength range from 400 to 700 nm.Light with a wavelength shorter than 400 nm or longer than 700 nm was considered as unimportant for photosynthesis, due to its low quantum yield of CO 2 ...
Since 1893, when the word "photosynthesis" was first coined by Charles Reid Barnes and Conway MacMillan, our understanding of the elements and regulation of this complex process is far from being entirely understood. We aim to review the most relevant advances in photosynthesis research from the last few years and to provide a perspective ...
Research about how C 4 photosynthesis ... which may be associated with the starch accumulation and sugar feedback inhibition of photosynthesis. High-light ... The datasets analyzed in this paper ...
Journal of Integrative Agriculture 2021, 20(1): 4â€"23 REVIEW Available online at www.sciencedirect.com ScienceDirect Crop photosynthetic response to light quality and light intensity Iram SHAFIQ1, 2*, Sajad HUSSAIN1, 2*, Muhammad Ali RAZA1, 2, Nasir IQBAL1, 2, Muhammad Ahsan ASGHAR1, 2, Ali RAZA1, 2, FAN Yuan-fang1, 2, Maryam MUMTAZ1, Muhammad SHOAIB3, Muhammad ANSAR4, Abdul MANAF4, YANG ...
Overall, the present study provides highly reliable ranges of photosynthetic pigment contents based on a significant number of papers and a wide diversity of species, which may supply reference ranges for future research on photosynthesis and plant acclimation. Particularly stable and useful for this purpose are Neo : Chl and the Chla : Chlb ...
It is the biochemical process that sustains the biosphere as the basis for the food chain. The oxygen produced as a by-product of photosynthesis allowed the formation of the ozone layer, the evolution of aerobic respiration and thus complex multicellular life. Oxygenic photosynthesis involves the conversion of water and CO 2 into complex ...
Many snow and ice models for Mars 14,44,45,46,47 use a broadband approximation to calculate the fluxes of solar radiation at depth. As an illustrative example, Fig. 1 compares the Greenland ...
This new paper suggests that dusty ice lets in enough light for photosynthesis to occur as deep as 9 feet (3 meters) below the surface. In this scenario, the upper layers of ice prevent the ...
Photosynthetic Research in Plant Science. Photosynthesis is a highly regulated, multistep process. It encompasses the harvest of solar energy, transfer of excitation energy, energy conversion, electron transfer from water to NADP +, ATP generation and a series of enzymatic reactions that assimilate carbon dioxide and synthesize carbohydrate ...