© Springer-Verlag GmbH Germany, part of Springer Nature 2019
E.-D. Schulze et al.Plant Ecologyhttps://doi.org/10.1007/978-3-662-56233-8_3

3. Light

Ernst-Detlef Schulze1 , Erwin Beck2, Nina Buchmann3, Stephan Clemens2, Klaus Müller-Hohenstein4 and Michael Scherer-Lorenzen5
(1)
Max Planck Institute for Biogeochemistry, Jena, Germany
(2)
Department of Plant Physiology, University of Bayreuth, Bayreuth, Germany
(3)
Department of Environmental Systems Science, ETH Zurich, Zurich, Switzerland
(4)
Department of Biogeography, University of Bayreuth, Bayreuth, Germany
(5)
Chair of Geobotany, Faculty of Biology, University of Freiburg, Freiburg, Germany
 
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Figa_HTML.png

Wide spatial and temporal gradients of light intensity characterise the light climate of a tropical mountain forest in the Andes of Ecuador. On average, less than 10% of the incoming radiation reaches the vegetation on the forest floor. However, spatially limited sunflecks can increase the light intensity several hundredfold for a short while, challenging the adaptability of the photosynthetic apparatus of the leaves. Note the dense cover of epiphytes on the branches of the trees. (Photo: M. Richter, Erlangen)

3.1 The Dual Significance of Light

Sunlight is by far the most dominant energy source for all life on Earth. For plants, as sessile photoautotrophic organisms, it is one of the most important environmental factors. Light varies in intensity and spectral composition from place to place, from season to season, and in the course of the daily photoperiod. In order to optimally harvest the energy of light and to minimise stress arising from insufficient or supraoptimal absorption of photons, plants have evolved multiple ways to modulate light exposure and photosynthesis (Scholes et al. 2011; Chaps. 9 and 12).

Furthermore, light quality (colour), intensity (quantum flux) and duration (day length) have a second function as environmental cues, which guide the plant through its entire life cycle. Developmental plasticity of plants enables effective adaptation and stress avoidance strategies and is to a large extent based on the ability to perceive and transduce light signals. Key transitions from seed to germination, through subsequent phases of vegetative growth, to flowering and finally senescence of the entire plant or its organs are regulated in response to a fluctuating light environment (Chap. 2, Sect. 2.​1.​4).

Plants have evolved photosynthetic pigments such as chlorophylls to harvest the energy of light, and a diverse set of photoreceptors to monitor the intensity, spectral composition and direction of light. Photoreceptors differ from the photosynthetic pigments with respect to absorption spectra, intracellular localisation and the effects of excitation by photons. While the light-harvesting systems are found in the chloroplasts/thylakoids, the sensors regulating developmental processes are localised in the cytosol and the nucleus. Excitation of chlorophyll initiates an electron transfer chain coupled to the production of adenosine triphosphate (ATP) and reduced nicotinamide adenine dinucleotide phosphate (NADPH). Excitation of photoreceptors triggers signal transduction processes and concomitant changes in gene expression. In spite of the fundamental differences in the molecular, biochemical and biophysical processes involved, the two functions of light for plants—as an energy source and as an environmental cue—are tightly interlaced. Many of the responses of a plant to changing light conditions result in an optimisation of utilisation of light for photosynthesis.

Terrestrial plants, as well as the majority of algae, use blue and red light for photosynthesis. The respective accessory pigments are carotenoids and chlorophyll b (in some algae also other chlorophylls), while the reactive photopigments are dimers of chlorophyll a, termed P680 and P700, respectively. The bulk of chlorophyll a serves as a light-harvesting pigment. Cyanobacteria and red algae have no chlorophyll b but have additional accessory pigments, which allow them to inhabit aquatic biotopes with an altered composition of the visible spectrum. Phycobiliproteins, assembled in the phycobilisomes, allow these organisms to effectively use blue-green light for photosynthesis and thus to inhabit deeper water layers or muddy waters where green algae cannot harvest sufficient light for photosynthesis. Red light is absorbed in the top layers of a water body and blue light is scattered, resulting in the dominance of blue green at greater depth. Red algae and cyanobacteria are able to dynamically modify the proportions of the red-absorbing and blue-green-absorbing pigments in the phycobilisomes to acclimate to special light climates (MacIntyre et al. 2002) (Fig. 3.1).
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig1_HTML.png
Fig. 3.1

Cyanobacteria are photosynthesising prokaryotes in saline as well as fresh waters. a Whips of cyanobacteria accumulating close to the surface in the southern Baltic Sea. The brownish colour of these unicellular organisms does not support their common name, blue-green algae, as the typical phycobilins are often overlaid by carotenoids (Rastogi et al. 2010) (Nordic Microalgae c/o SMHI Oceanographic Unit Sven Källfelts gata 15, SE-426 71 Västra Frölunda Sweden; Copyright by Bengt Karlson). b Colourful zonation of cyanobacteria in the Morning Glory Pool at the Yellowstone National Park, Wyoming (from Encyclopaedia Britannica Online (2016); license no. SSTK-041E2-CF2B by Shutterstock)

A minor yet highly relevant component of the solar radiation reaching the Earth’s surface is ultraviolet (UV) light. Because of its high energy it can be damaging to the macromolecules of biological systems, including DNA. Therefore, plants—like other organisms colonising the land—have evolved very effective mechanisms for protection against UV radiation and to repair UV-induced damage. Activation of repair is in part dependent on the perception of UV light as a signal.

Avoidance of damage by excessive irradiation or of low light, as well as acclimation to different and changing light climates, occur at different levels: morphological, anatomical (Chaps. 9 and 12), cellular, subcellular and molecular. Several of the responses, particularly the morphological and structural ones, are principally irreversible—for instance, the formation of shade and sun leaves. Others are dynamic and reversible, such as hinge movements of entire leaves or the displacement of chloroplasts within the cells. In a few cases (e.g. the shade avoidance response and state transitions), mechanisms have been elucidated in molecular detail.

This chapter will focus on avoidance and acclimation responses of plants to fluctuating light climate and to UV exposure, as well as the perception systems underlying light-controlled plasticity of plant development. For the mechanisms of photosynthesis and comprehensive accounts of light signal transduction, the reader is referred to plant physiology and plant biochemistry textbooks (e.g. Buchanan et al. (2015) and Taiz et al. (2015)).

3.2 Visible Light

3.2.1 Avoidance of Light Stress and Permanent or Dynamic Acclimation

Lack of light, interspersed with high light intensity of short-term sunflecks, combined with a change in the spectral composition of the radiation (Fig. 3.2), is a major stressor of the vegetation on the forest floor and, less dramatically, in the shade crown of a tree (Chap. 9). As the leaves in the canopy absorb mainly blue and red light, green light is enriched in the spectral composition of the subcanopy space (“green shade”). Not visible to the human eye is another change in the spectrum of sunlight by passage through the forest canopy: the dramatic decrease in the red to far-red ratio, as the leaves absorb red and blue light but transmit and reflect far-red light (Chap. 9).
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig2_HTML.jpg
Fig. 3.2

Seasonal change of the light climate in an oak–pine forest. a Daily courses of the light intensity above the canopy. b Light attenuation in the New Jersey Pine Barrens oak-pine forest in early February (left side) and at the end of July (right side). Note the different scales used for the colour bars for winter and summer. All units are in micromoles of photons per square metre per second. (Modified from Schäfer and Dirk (2011))

Responses to low light are obvious at the forest edges where subcanopy plants (e.g. ferns and other herbaceous species) show a pronounced phototropic reaction—that is, they grow towards the higher radiation intensity outside the forest. Plant life forms that avoid the attenuated light climate under the canopy of a forest are functional types such as epiphytes and lianas; both are light parasites that use trees as support to reach more favourable light conditions (Chap. 9).

On the other hand, a surplus of light intensity can also cause problems, as photosynthetic pigments cannot avoid energy absorption, which may overstrain the photosynthetic capacity of the chloroplasts and produce reactive oxygen species (ROS) (Chap. 2, Sect. 2.​2.​3). Leaves often avoid excessive irradiation by adopting a parallel position to the incident light. The hanging leaves of Eucalyptus are well known from the “shadeless forests” in Australia. Likewise, the upright position of the leaves of the characteristic giant rosette plants of the tropical high mountains (Fig. 4.​25) has been described as a mechanism for avoiding direct solar radiation.

The position of the leaf lamina of many plants varies during the daily light period, changing the angle and intensity of the incoming radiation. Such leaves often have joints between the petiole and the lamina, known as pulvini (Fabaceae, Oxalidaceae), which enable fast changes in the position of the entire leaves or at least the pinnules (Chap. 9). The reaction of the North American wood sorrel (Oxalis oregano) to short-term high-light stress from a sunfleck is shown in Fig. 3.3.
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig3_HTML.png
Fig. 3.3

Avoidance of light stress by a change in the leaf angle. Oxalis oregano, the wood sorrel, is a particular shade plant whose leaves avoid light stress of bright sunflecks by rapidly changing the angle of their leaflets by virtue of pulvini. When the sunfleck is over they slowly return to their horizontal position, which is optimal for light harvesting. (After Björkman and Powles (1981))

Dynamic acclimation to the light intensity can be observed at the cellular level too. In many algae and mosses, but also in vascular plants, chloroplasts can change their position in the cell. They accumulate on the light-exposed surfaces in the case of low light and at the light-parallel cell boundaries when the radiation becomes too strong (Fig. 3.4). Such readily reversible intracellular reactions are mediated through the blue light receptor phototropin (Wada 2013).
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig4_HTML.jpg
Fig. 3.4

Chloroplast movement is important for avoiding photodamage under high light and for efficient light harvesting under low light. Object: Leaflet of the moss Funaria hygrometrica, consisting of a single layer of chloroplast-containing cells. a Position on the lateral walls of the cells in high light. b Position of the chloroplast on the light-exposed side in low light (Nultsch 2001). A special protein named Chloroplast Unusual Position 1 (CHUP 1) is unique in positioning and moving chloroplasts in the Arabidopsis thaliana leaf cell. It is assumed that CHUP 1 connects the chloroplast with the actin cell skeleton and the cell membrane (Oikawa et al. 2003, 2008; Usami et al. 2012)

Besides such short-term acclimations (with a reaction time in minutes), development of leaves responds to the light environment by the so-called shade avoidance syndrome and the formation of sun and shade leaves. The shade avoidance syndrome is part of the morphogenetic reaction of seedlings when they are grown under low light intensities caused, for instance, by competing neighbours. Owing mostly to the identification of the main photoreceptors in the model plant Arabidopsis thaliana and the availability of respective mutants, knowledge on the molecular regulation of plant growth and development by light is rapidly accumulating (Casal 2013; Lau and Deng 2012; Liu et al. 2011; Lau and Deng 2012; Kami et al. 2010). Because a separation of acclimation and development is principally difficult here, we focus on the reactions triggered by light stress—that is, by either insufficient or supraoptimal light intensities.

The typical shade avoidance reaction in A. thaliana can serve as an example illustrating the integration of light cues and growth at the molecular level. Seedlings show an inhibition of lamina expansion, while petiole elongation and stem growth are enhanced. Phytochromes (PhyB) and cryptochromes (Cry1 and Cry2), as photoreceptors, have been demonstrated to play specific roles in these differential growth responses. Phytochromes especially are ideally suited to convey information on shade occurrence because of their switching between red and far-red absorption and the pronounced difference in the ratio of these two light qualities between sunlight and shade light. Low red to far-red ratios typical of shade light reduce the levels of active PhyB. Because PhyB inactivates a class of transcription factors named phytochrome-interacting factors (PIFs), their activity increases under shade light conditions. PIFs have been demonstrated to induce growth responses—for instance, through the activation of genes involved in cell wall biosynthesis. In addition, PIFs influence phytohormone signalling. They enhance auxin concentrations through the activation of auxin biosynthesis. Promotion of shade avoidance reactions by the other growth hormones gibberellic acid and brassinosteroids can be attributed to a positive effect on PIF abundance. The function of cryptochromes in shade avoidance is associated with degradation of the constitutive photomorphogenesis protein 1 (COP1) (Fig. 3.22), which is jointly triggered by cryptochromes and phytochromes, albeit in an unknown fashion different from photomorphogenesis (Casal 2013).

The shade avoidance reaction as observed under low red to far-red illumination must be attenuated when exposure to this light climate is prolonged and also when the leaf reaches into a space with sufficient light intensity. Similarly, sunflecks reduce the shade avoidance response. An autoregulatory mechanism involving transcription factors such as HY5 (Figs. 3.21 and 3.22) has been described that helps to avoid exaggerated shade avoidance responses (Jiao et al. 2007).

Through the shade avoidance reaction the plant strives to escape unfavourable low light intensity, using its resources to grow into a better-illuminated space. A typical example is the ground flora in a forest, consisting of herbs and bushes with elongated internodes and smaller leaves (Chap. 9). The formation of sun and shade leaves, on the other hand—which is well known from tree crowns—can be understood as differential acclimation of a plant’s foliage to high light intensities in the outer part (the so-called sun crown) and low light intensity in the centre of the crown (termed the shade crown). Such differentiation within the same shoot is possible because of the modularity of the plant structure, which is composed principally of similar units called phytomers, whose development can be individually guided by local environmental cues (Chap. 2, Sect. 2.​4.​3). Sun leaves deliver the major share of the photosynthetic gain owing to an acclimated photosynthetic machinery but also exhibit greater dynamics of cell respiration. In shade leaves, both processes run at lower rates, which shows their autonomy and ensures survival. Export of assimilates from shade leaves to other plant organs, however, may be low or non-existent. Further characteristics of thick sun leaves and thin shade leaves are summarised in Table 3.1.
Table 3.1

Comparison of characteristic traits of shade- and sun-acclimated leaves

Characteristic

Sun leaf

Shade leaf

Structural traits

   

Dry mass per area

High

Low

Leaf thickness

Thick

Thin

Palisade parenchyma thickness

Thick

Thin

Spongy parenchyma thickness

Similar

Similar

Density of stomata

High

Low

Chloroplasts per area

Many

Few

Thylakoid density in the stroma

Low

High

Thylakoids per granum

Few

Many

Biochemical traits

   

Chlorophyll per chloroplast

Low

High

Chlorophyll per area

Similar

Similar

Chlorophyll per dry mass

Low

High

Chlorophyll a to chlorophyll b ratio

High

Low

Light-harvesting complexes per area

Few

Many

Electron transport components per area

High

Low

ATPase per area

High

Low

RubisCO per area

High

Low

Nitrogen per area

High

Low

Xanthophylls per area

High

Low

Gas exchange

   

Photosynthetic capacity per area

High

Low

Dark respiration per area

High

Low

Photosynthetic capacity per dry mass

Similar

Similar

Dark respiration per dry mass

Similar

Similar

Carboxylation capacity per area

High

Low

Electron transport capacity per area

High

Low

Quantum yield

Similar

Similar

Light response curve

images/72100_2_En_3_Chapter/72100_2_En_3_Figb_HTML.gif

images/72100_2_En_3_Chapter/72100_2_En_3_Figc_HTML.gif

RubisCO ribulose-1,5-bisphosphate carboxylase/oxygenase. (Modified from Lambers et al. (2008))

In contrast to trees, herbs are amenable to experimental approaches aimed at understanding the formation of physiologically and anatomically different types of leaves (Fig. 3.5). The signals indicating a high-light climate are perceived by the mature leaves, which trigger the respective development of the young leaves in the apical shoot meristem via long-distance signalling. Interestingly, the ambient CO 2 concentration has effects on the formation of sun and shade leaves similar to those of the light intensity. High CO2 leads to the formation of sun leaves (Prior et al. 2004; Pritchard et al. 1999).
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig5_HTML.png
Fig. 3.5

Differential development of sun and shade leaves in Chenopodium album in response to light intensity and CO2 concentration. Sun leaves have a dense mesophyll with at least two layers of palisade parenchyma cells, which originate from a directional change of the cell divisions. The scale bars indicate 100 μm (Kim et al. 2005)

From experiments such as those shown in Fig. 3.5 it was concluded that, at least in dicots, systemic signalling (Chap. 2, Sect. 2.​2.​4) depending on the light environment influences the development of new foliage and thus triggers acclimation to the actual light quality and intensity. Long-distance signalling from mature to developing leaves is also known from herbivore and pathogen attacks (Chap. 8), and most likely a multitude of signals is involved in mediating different types of responses. Acclimation of the photosynthetic apparatus appears to be under the control of the redox state. Thus, redox signalling may also be part of the long-distance conveyance of biochemical information. Also, the concentrations of photosynthates may act as an indirect cue for the light climate of mature leaves. However, since the formation of sun and shade leaves is a highly complex syndrome, involvement of various further signals can be expected.

At the cellular level the epidermal cells, in particular, contribute to the protection of the mesophyll cells against high-light stress. A felt-like layer of dense hairs, which reflects incident light, is found on young leaves of many plant species. Although it can be kept for permanent protection, it mostly disappears during maturation of the leaves (Chap. 9). Protective pigments, which absorb predominantly shortwave radiation, are also found in the epidermis of young leaves. A typical example is the so-called juvenile anthocyanin (Fig. 3.25), which protects not yet fully green leaves during the development of the photosynthetic apparatus. Later, protection is conferred mostly by carotenoids in the antennae. This is apparent from the disappearance of the pigments in the epidermis.

3.2.1.1 Ultrastructural Acclimation to the Light Environment

Chloroplasts of sun and shade leaves differ greatly with respect to their thylakoid systems (Fig. 3.6). Chloroplasts of sun leaves possess only thin grana, while chloroplasts of shade leaves show large grana stacks. The importance of the size of the thylakoid system in a chloroplast becomes obvious when the molecular structure of the thylakoid membranes and the functions of the photosynthetic protein complexes are considered. As photosystem II (PS II) has a larger antenna than photosystem I (PS I), over half of the total chlorophyll a, almost all of the chlorophyll b and most of the carotenoids (β-carotene and xanthophylls) are associated with PS II. The largest part of a cross-section through a photosystem consists of its antennae. PS II and its antennae are located in the so-called appressed regions—the contact zones of the stacked thylakoids. The number of antenna complexes—especially the outer or mobile antennae, which feed excitation energy to the photosynthetic reaction centre—is variable. A shade chloroplast contains a very large number of light-harvesting systems relative to the reaction centres, owing to the large thylakoid stacks and big antennae. This corresponds to a lower chlorophyll a to chlorophyll b ratio, indicating a higher proportion of the light-harvesting chlorophyll b–containing antennae around photosystem II (Table 3.1) (Kitajima and Hogan 2003). Chloroplasts of sun leaves, in contrast, contain smaller antennae and thus the ratio of antennae to reaction centres is smaller. The different organisation of the thylakoid systems of sun leaf and shade leaf chloroplasts can be considered a physiological acclimation to the ambient light environment, as both combinations allow optimal utilisation of the incident light. Such differences are observed not only in sun and shade leaves but also between the upper and lower mesophyll cells of dorsiventral leaves. Chloroplasts of the better-illuminated upper side appear as sun leaf chloroplasts, while those of the lower side (usually the spongy parenchyma) show characteristics of shade leaf chloroplasts. As the thylakoid system of a chloroplast is a dynamic substructure, fixing a leaf upside down accordingly results in a reorganisation of the thylakoid systems of the mesophyll cells.
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig6_HTML.png
Fig. 3.6

Acclimation of chloroplast ultrastructure to the light environment. a Chloroplast from a sun leaf of tobacco. The small grana stacks are typical of chloroplasts from high-light leaves (Hall and Rao 1994). b Transverse section of a chloroplast from a shade leaf of snapdragon (Antirrhinum majus), showing enhanced thylakoid stacking (Strasburger 1983)

The range of light energies known to be used for photosynthesis is huge. A champion of using low photosynthetic light intensities is a green sulphur bacterium (Chlorobium phaeobacteroides) from the Black Sea, collected from an 80-m depth, where the light intensity is 3–10 nmol quanta m−2 s−1. Thus, a single bacterium receives no more than 300 photons per second, whereas 1016 photons per second impinge on a medium-sized leaf of a terrestrial plant. Owing to high concentrations of light-harvesting pigments and very low maintenance energy requirements, the sulphur bacterium can survive but takes years to double (Overmann et al. 1992). In spite of its acclimation to the low-light environment, the efficiency of energy transfer in its photosynthetic apparatus (chlorosomes) is no more than about 60% of the absorbed radiation. Even fast-growing terrestrial plants with optimised photosynthetic membranes use less than 50% of the absorbed photosynthetic active radiation for photosynthetic CO2 assimilation (Scholes et al. 2011) (Fig. 3.7).
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig7_HTML.png
Fig. 3.7

Gradients of chlorophyll, light absorption and photosynthetic activity across a spinach leaf. a Thin cross-section (Modified from Munns et al. (2010)). b Profiles of light intensity, absorbed light, chlorophyll and photosynthetic 14CO2 fixation across a spinach leaf. CO2 fixation follows best the profile of absorbed light (based on Nishio et al. (1993) and Evans (1995))

On average, a dorsiventral leaf (e.g. of spinach) absorbs about 85% of the incident light between 400 and 700 nm. This is mainly due to extension of the light path in the leaf by scattering; depending on the epicuticular fine structure, up to 10% is reflected and the remaining (~5%) is transmitted (Munns et al. 2010) (Chap. 9).

3.2.2 Overexcitation and Damage to Photosynthetic Membranes

Radicals generated by overexcitation of PS II (Chap. 2, Sect. 2.​2) result first in photoinhibition (destruction of the D1 protein of the affected reaction centre). Prolonged photoinhibition leads to greater damage and finally to the destruction of the photosynthetic membranes. The potential for radical formation and photodamage is even enhanced by the oxygen-rich micro-environment of the chloroplast and the strong oxidising conditions needed for the crucial reaction of photosynthesis—the oxidation of water.

3.2.3 Flexible Acclimation to Changes in Light Intensity

The photosynthetic apparatus of higher plants serves two seemingly opposing functions: (a) harvesting of light and transfer of the excitation energy to the reaction centre; and (b) dissipation of excessively absorbed light harmlessly as heat in order to prevent photodestruction of the thylakoids. Not only can the ratio of the antennae to the reaction centres vary but also the dimensions of the antenna complexes can change in response to the light intensity (Fig. 3.8).
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig8_HTML.png
Fig. 3.8

Long- and short-term acclimation of photosynthetic membranes in response to the light environment and to a heat shock. As long-term acclimation under high-light conditions the antennae of photosystem II (PS II) are smaller than those under low-light conditions, thereby reducing light absorption. Short-term regulation: overexcitation of PS II due to excessive light triggers the state transition, a rapid change in the antennae sizes of PS I and PS II. Light-harvesting complex II (LHC II) becomes phosphorylated and partly associates with PS I instead of PS II. (Modified from Anderson and Andersson (1988))

The so-called major antenna of PS II, LHC II (light-harvesting complex II, usually organised as trimeric complexes), together with the minor peripheral light-harvesting complexes CP26 and CP29 (chlorophyll proteins, named after their molecular weights in kDa), can dissociate from the core antenna complexes, which are also chlorophyll-containing proteins (CP43 and CP47) surrounding the reaction centre of PS II. When the energy pressure on the reaction centre is high and PS II is overexcited relative to PS I, the so-called state transition occurs (Fig. 3.8). The plastoquinone pool becomes over-reduced. This activates protein kinases that phosphorylate threonine residues in the peripheral antenna proteins. Phosphorylation leads to an accumulation of negative charges and a dissociation of the peripheral antennae from the core antennae. At the same time, the connection of the appressed regions loosens, allowing lateral movement of the peripheral antennae and thus a diminution of the PS II supercomplex (Minagawa 2013). Simultaneously, the extent (as well as the efficiency) of light harvesting of PS I increases, because part of the peripheral antennae of PS II can associate with PS I, thereby balancing the excitation of both photosystems (Fig. 3.9). Such a balance is essential for optimal utilisation of the light energy and is likewise important for the avoidance of PS II overexcitation (over-reduction of the pools of its redox compounds such as Q B).
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig9_HTML.png
Fig. 3.9

Linear photosynthetic electron transport (from water to nicotinamide adenine dinucleotide phosphate (NADP+)) in the context of utilisation of absorbed energy, and dissipation of excess energy as fluorescence and heat. Acclimation of the photosynthetic machinery to high excitation by the so-called state I–state II transition (dissociation of the light-harvesting complex (LHC) from photosystem II (PS II)) is also shown. The majority of the absorbed light energy is dissipated as heat, and a small proportion (<5%) is emitted as fluorescence. A variable proportion can be used for photochemical work (oxidation of water and electron transport). Under strong illumination the peripheral antennae of PS II can dissociate from the photosystem and at least in part associate with PS I (state transition; Fig. 3.8). In this way, PS II absorbs less and PS I absorbs more light energy. Since the entire system is dynamic, the rate constants (thickness of arrows) can change—for example, by overexcitation. Photochemistry: presented are the constituents of the linear electron transport (After Schreiber et al. (1994))

Even in the balanced electron flow, light intensity frequently exceeds the capacity of its utilisation for photosynthesis. According to theoretical considerations, 8 mole quanta are required for assimilation of 1 mole of CO2. However, because of the way in which ATP synthesis is coupled to the linear electron flow, slightly more quanta are necessary. The highest measured quantum efficiency (lowest quantum requirement) is 9.4 mole quanta per mole of assimilated CO2. For such measurements, the light intensity must be limiting, as is the case in the linear range of the light response curve (Fig. 3.10, curve a).
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig10_HTML.png
Fig. 3.10

Interpretation of a light response curve of photosynthetic CO2 uptake. Curve a: linear increase of the photosynthetic rate in low light (limitation of photosynthesis by light intensity). Curve b: measured photosynthetic rate indicating light saturation at high photon flux density (PFD). Curve c: calculated proportion of reduced (closed) photosystem II (PS II) reaction centres with excessive light. Curve d: Excess light energy (corresponding to the horizontal arrows between curves a and b). (Modified from Björkman and Demming-Adams (1994))

When the light response curve of CO2 uptake deviates from the linear relationship (Chap. 12), more light is absorbed than can be used for photosynthesis and the plant is confronted with the need for energy dissipation. At the light intensity of a sunny day (800–1000 μmol quanta m−2 s−1), already 500–600 μmol quanta m−2 s−1 are in excess. The problem of de-energisation of the photosynthetic membranes is aggravated when water shortage necessitates closure of stomata and concomitant lowering of the intercellular CO2 concentration. Similarly, on a bright day in winter, stress arises when the low ambient temperatures greatly decelerate metabolism and metabolite fluxes while high radiation intensities impinge on the leaves. Plants have evolved several mechanisms to cope with this challenge.

In Fig. 3.10, curve d shows the excessive photon flux density that needs to be dissipated. Less than 5% of the absorbed energy can be dissipated as chlorophyll a fluorescence (Fig. 3.9), which is complementary to the portion that can be utilised for photochemistry (photosynthesis). Because of this relation, photosynthesis can be followed indirectly by measurement of chlorophyll a fluorescence. This is illustrated in Fig. 3.11.
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig11_HTML.png
Fig. 3.11

Analysis of chlorophyll a fluorescence from photosystem II (PS II) of a pre-darkened leaf. The first peak reflects the time course of chlorophyll a fluorescence intensity triggered by a single saturating flash on a pre-darkened leaf. A series of saturating flashes results in a decrease of fluorescence by two mechanisms: q P is the quench of fluorescence by photosynthetic electron flow and reoxidation of the reaction centres, while q E reflects a quench resulting from an increasing rate of energy dissipation in the antennae of PS II. F 0 is understood to result from fluorescence of some antenna chlorophylls, while F V is the fluorescence that is complementary to photosynthetic electron flow. Actinic light (low intensity) is necessary to maintain the photosystems in an activated state. (After Schreiber et al. (1986))

The intensity of fluorescence is composed of a small contribution from the antennae (termed F 0) and a major component from chlorophyll a of the reaction centre. The latter is termed variable fluorescence (F V). When the photosystem is completely reduced—that is, when there is no acceptor of electrons available—F V is maximal (F max = F 0 + F V). A return of F V to the ground state (F 0) takes about 30 s. A series of subsequent saturating light flashes decreases F m to F m′ because of a quench of the fluorescence of the fully reduced (“closed”) reaction centres. This quench is again composed of several components: q p is the portion of fluorescence that is quenched by photochemistry (i.e. photosynthesis); q E is the so-called non-photochemical quenching (NPQ) or energy-dependent quenching, which depends on the trans-thylakoid pH gradient, on the concentrations of the xanthophyll zeaxanthin and on the antenna-associated protein PsbS. NPQ is the major mechanism of photoprotection. Excess light energy absorbed by the antennae of PS II is thermally dissipated. As Fig. 3.11 shows, this quench increases with increasing saturation/reduction of PS II.

The mechanism of NPQ is complicated insofar that the same ensemble of compounds has to mediate antagonistic reactions—namely, feeding of excitation energy to the reaction centre (of PS II) as well as dissipation of excess light energy by de-energisation of excited chlorophyll molecules. Conformational changes by protonation of the involved proteins, especially of PsbS by an acidic thylakoid lumen pH, result in switching from excitation to dissipation mode (Ahn et al. 2008; Correa-Galvis et al. 2016; Fan et al. 2015). The dissipative conformation is stabilised by zeaxanthin, which is a product of the xanthophyll cycle (Fig. 3.12). In high light and correspondingly low thylakoid lumen pH, violaxanthin de-epoxidase is activated, which catalyses the formation of zeaxanthin from violaxanthin. A model has been proposed that integrates the fast component of NPQ via protonation of PsbS and the slow component, the formation of zeaxanthin by the de-epoxidation of violaxanthin and subsequent association with PsbS (Zaks et al. 2012, Fig. 3.13). The fast component may be restricted to the already detached antennae, whereas the slow component appears to happen in the core antennae, including the minor chlorophyll proteins, and thus protects PS II from overexcitation (Holzwarth et al. 2009). The quantitatively dominant xanthophyll of the thylakoid membranes, lutein, can react in a way similar to zeaxanthin (Li et al. 2009). A lutein-5,6 epoxide cycle has been reported, which is driven by light-activated violaxanthin de-epoxidase and zeaxanthin epoxidase under low light (Matsubara et al. 2008). The physiological importance of the xanthophyll cycles is underlined by the increase and decrease in the concentrations of their components upon transfer of a plant from low light to high light, and vice versa (Fig. 3.14). Also, one of the first common garden experiments with A. thaliana wild-type and mutant plants demonstrated that plants with defects in NPQ suffer a significant loss of fitness (determined as seed yield) under naturally fluctuating light conditions (Kühlheim et al. 2002).
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig12_HTML.png
Fig. 3.12

The xanthophyll cycle in chloroplasts. In high light and an acidic luminal pH of the thylakoids, violaxanthin is converted into zeaxanthin by the de-epoxidase; in darkness and a slightly alkaline luminal pH, reoxidation of zeaxanthin to violaxanthin is catalysed by the epoxidase (Heldt and Piechulla 2010)

/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig13_HTML.png
Fig. 3.13

Model for the switching of the photosystem II (PS II) antenna from light harvesting into the energy dissipation mode by non-photochemical quenching (NPQ). a Excitation and flow of energy in PS II under moderate light. The protein PsbS is inactive; the xanthophyll pool consists mainly of violaxanthin (VX) with a very small amount of zeaxanthin (ZX). b Excitation of PS II with high irradiance creates a high proton motive force (ΔpH and ΔΨ). Acidic pH activates (protonates) PsbS, which associates with the antenna, leading to an attenuation of energy transfer to the active centre of PS II and higher dissipation of energy as heat. The Chla*–PsbS* complex is stabilised by zeaxanthin, which is produced from violaxanthin by the pH-activated violaxanthin de-epoxidase (VDE). In low light, NPQ is small or even almost absent and the energy is used for photosynthesis (harvesting mode)

/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig14_HTML.png
Fig. 3.14

Dynamic changes in xanthophyll pools of young cotton leaves upon changes of the light intensity from low to high a, and vice versa b. In contrast to the xanthophyll neoxanthin, for which no cycle is known, the xanthophylls undergoing de-epoxidation and epoxidation, respectively, respond readily to a change in the light environment. (Modified from Björkman and Demming-Adams (1994))

At the metabolic level of the Calvin cycle, even under conditions of closed stomata, a possibility to avoid or at least reduce overexcitation arises from the oxidative photosynthetic carbon cycle (photorespiration) which, under low CO2 (due to decreased conductance of the stomata) and high light, releases CO2 from glycine decarboxylation. This internal CO2 keeps the photosynthetic electron flow running through consumption of NADPH and ATP in the Calvin cycle (for the reactions and compartmentation of the oxidative carbon cycle, see plant biochemistry textbooks). Likewise, photosynthetic reduction of nitrite or sulphate and the subsequent formation of amino acids require these photosynthetic primary products. However, the rates of the latter pathways are comparatively small. Oxygen may also be reduced photosynthetically in the so-called Mehler reaction (Chap. 2, Sect. 2.​2), giving rise to ROS, which can be detoxified by a sequence of reactions, finally resulting in the consumption of NADPH (the “waterwater cycle” or pseudocyclic photosynthetic electron flow) (Fig. 3.15).
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig15_HTML.png
Fig. 3.15

The so-called water–water cycle or pseudocyclic photosynthetic electron transport. This occurs as a recourse of the electron flow when the chloroplastic pools of nicotinamide adenine dinucleotide phosphate (NADP+) and nicotinamide adenine dinucleotide (NAD+) are depleted because of an over-reducing environment and no electron acceptor other than oxygen is available. Reduction of oxygen produces the oxygen anion (superoxide). The oxygen anion dissociates to oxygen and peroxide, which is reduced to oxygen and water by ascorbate peroxidase. (For more details on the responsible enzymes, Chap. 2, Sect. 2.​2)

Box 3.1: A Day in the Life of a Tree Leaf: Dynamic Acclimation of Leaf Performance to Short-Term Environmental Changes

The effect of sunflecks on the photosynthetic performance of leaves (e.g. of a tree crown) has been discussed controversially with respect to the question as to whether a frequent change in the light intensity enhances or decreases their photosynthetic efficiency. The basis of such considerations is the observation that once a leaf is in a photosynthetically active state, the increase in the rate of photosynthesis is fast (occurring in a few seconds), while the return to the “shade state” is in the range of minutes, resulting in overall elevated efficiency (Fig. 3.16c). On the other hand, such a consideration must also take note of the nonlinear quantum efficiency of varying light intensities, which is described by the light response curve of net CO2 uptake (Figs. 3.10 and 3.16a). Sunflecks from direct sunlight commonly surpass the range of the linear or nearly linear relation between light intensity and photosynthetic CO2 uptake. In the examples shown in Figs. 3.16a and 3.17a for the tropical gymnosperm Podocarpus falcatus, the increase in photosynthetic CO2 uptake was minimal beyond a photosynthetically active radiation (PAR) intensity of about 400 μmol quanta m−2 s−1. The sunflecks, however, reached up to 1500 μmol quanta m−2 s−1 (Fig. 3.17a). Therefore, the sunflecks were of lower photosynthetic quantum efficiency than low light. When the same amount of PAR was supplied to a Podocarpus leaf over the same time period as continuous radiation of low intensity (Fig. 3.16b) or as artificial sunflecks (termed lightflecks; Fig. 3.16c), the photosynthetic gain was obviously higher under continuous low light. Under natural conditions, not only the light intensity varies (Fig. 3.17) but also the conductivity (g s) of the stomata responding to the water status of the leaves (Chaps. 10 and 12). When stomatal conductivity is low, as it generally is in the afternoon, the photosynthetic efficiency of sunflecks is even lower because of a low internal CO2 concentration. This in turn indicates the importance of energy dissipation by the chloroplasts. The actual quantum efficiency of CO2 assimilation over an entire day was only 72% of the (theoretical) quantum efficiency of the same amount of quanta administered over the same time period, when provided as continuous low light (Table 3.2). This difference is, of course, not a constant. However, it indicates the extent to which the real light climate is less effective than artificial illumination. Continuous artificial low light mimics, to some extent, the diffuse radiation in the shade of a tree canopy which, for an entire forest canopy, is photosynthetically more efficient than direct irradiation (Mercado et al. 2009). In comparison with the mentioned minimal quantum requirement for photosynthetic CO2 assimilation of 9.2 μmol quanta per μmol CO2, the data in Table 3.2 suggest a sixfold lower quantum efficiency (54 mole quanta per mole of CO2). Under adverse environmental conditions the apparent quantum efficiency can further decrease tremendously—for example, when stomata are closed under drought and merely 1–2% of the incident visible light can be used for photosynthesis (Osmond et al. 1997). It is important to note that quantum efficiency has a different meaning in the ecological context than it does in the photosynthetic light reaction, where only those quanta that conduct photochemistry are counted. The difference is due to the energy of a high proportion of absorbed quanta being dissipated as heat under natural conditions.
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig16_HTML.png
Fig. 3.16

Photosynthetic CO2 net uptake by leaves of young Podocarpus trees (shade leaves) under the canopy of an old evergreen tropical forest. a Light response curve. The data show the mean values for 2–5 saplings per site with three repetitions per leaf (± standard error (S.E.)). bc Analysis of the sunfleck effect on photosynthetic CO2 uptake by two artificial light conditions, providing the same amount of photosynthetically active radiation (PAR) over an identical time span. b PAR provided as a constant photon flux density of 83 μmol m−2 s−1. c PAR provided as intermittent lightflecks lasting 30 s at an intensity of 200 or 400 μmol m−2 s−1 superimposed on a basic intensity of 40 μmol m−2 s−1. (Modified from Strobl et al. (2011))

/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig17_HTML.png
Fig. 3.17

A day under the canopy in the natural forest. Daily courses of a CO2 net uptake and ambient photosynthetically active radiation (PAR), and b stomatal conductance (g s) and transpiration (E) of young Podocarpus leaves and of air temperature on a sunny day (24 November 2006). (Modified from Strobl et al. (2011))

Table 3.2

Effect of sunflecks on the photosynthetic gain of a tree. Photosynthetic efficiency of the subcanopy light climate compared with virtual constant illumination of the same PAR magnitude applied over the same time period to leaves of Podocarpus falcatus (Modified from Strobl et al. (2011))

Daily sum of PAR [mol m−2 day−1]

Calculated PARav over the day (μmol m−2 s−1)

Theoretical daily CO2 net uptake on the basis of PARav (mmol m−2 day−1)

Measured daily CO2 net uptake (mmol m−2 day−1)

Apparent efficiency of actual PAR (%)

1.5

50.5

38.3

27.7

72

PAR photosynthetically active radiation, PAR av average PAR intensity

3.2.4 Continuous Light

In greenhouses, plants are frequently cultivated under continuous light for enhancement of biomass production. On the other hand, continuous light can induce severe injury in many plants. Plants sensitive to continuous light (e.g. eggplant, peanut, some cultivars of potato and many cultivars of tomato, but also some lichens and mosses) react with lower rates of photosynthesis, leaf chlorosis and necrosis (Velez-Ramirez et al. 2011). Several reasons have been identified for that damage. Some of it results from a disturbance in photosynthetic carbohydrate metabolism. Assimilatory starch accumulates during the daily light period between the grana of the chloroplasts. Suppression of nocturnal starch degradation leads to continuous deposition of starch and finally destruction of the chloroplast ultrastructure. Enhanced formation of radicals and ROS are other reasons, in particular when the applied light contains a high proportion of blue light. Usually the natural day/night cycle is accompanied by an oscillation in the temperature. Continuous light in combination with a constant temperature can disturb the circadian clock and affect the development of the plant. Is continuous light not a natural phenomenon of the polar summer beyond the polar circles, one might ask? By no means! Even in the polar summer, the light intensity (and light spectral distribution) oscillates in the daily rhythm concomitantly with the temperature. Thus, artificial continuous light (and temperature) provided in a greenhouse represents an environment without an analogy in nature. It is therefore maybe more surprising that some plants can grow well in continuous light than that many plants cannot.

3.2.5 Light Triggers Plant Adaptation and Acclimation to the Environment

3.2.5.1 Photoreceptors

Because of its role as an energy source, light is arguably the most important environmental factor for a plant. Consequently, perception of light and translation of the respective signals control many aspects of plant development and are crucial for adaptation to diverse habitats. The sensing of light conditions is enabled by a range of receptors, which are specific for particular wavelengths of the visible spectrum (Fig. 3.18). Phytochromes are receptors for red and far-red light, cryptochromes and phototropins are receptors for blue light. The photoreceptors for red and blue light principally consist of a protein and a chromophore. The chromophores are molecules absorbing light of the respective wavelengths (for example, flavin adenine dinucleotide (FAD) and tetrapyrroles). During evolution they have been recruited for a wide range of biological functions. Upon photon absorption the chromophores undergo conformational changes that are transmitted to the protein and thereby change its activity. For UV light a special photoreceptor has recently been identified (UVR8) which, because of the UV absorption by aromatic amino acid residues, does not require a chromophore.
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig18_HTML.png
Fig. 3.18

Spectrum of visible light and the respective photoreceptors of terrestrial plants. Perception of shade light signals by photosensory receptors. a Spectral photon distribution of sunlight and shade light. b Impact of different wavebands on the status of phytochromes, cryptochromes, phototropins and UVR8 (Casal 2013)

Phytochromes
Phytochromes are proteins with an open-chain tetrapyrrole as a chromophore that changes its configuration upon absorption of red or far-red light. The effective ratio of red to far-red, ζ, refers to the broad spectral peaks of phytochrome in the red and far-red and can be calculated by Eq. 3.1 (Chelle et al. 2007):
 $$ \zeta =\frac{E_{\mathrm{R}}}{E_{\mathrm{FR}}}=\frac{E_{\mathrm{R}\ \mathrm{direct}}+{E}_{\mathrm{R}\ \mathrm{diffuse}}}{E_{\mathrm{FR}\ \mathrm{direct}}+{E}_{\mathrm{FR}\ \mathrm{diffuse}}} $$
(3.1)

where E R and E FR are the spectral irradiances from the sun and the sky and the complex radiative transfer reactions within the canopy. E R comprises the radiation between 655 and 665 nm and accordingly E FR comprises that between 725 and 735 nm. Whereas ζ of the sunlight is around 1.2, it decreases in the understorey space to less than 0.2. Reactions that are triggered by (red) light (e.g. light-dependent seed germination) thus do not take place in the shade of a dense canopy. The situation changes in forest gaps where, because of the direct irradiation, ζ is high and seeds of pioneers can germinate rapidly (Chap. 2, Sect. 2.​4). In contrast, the time span of sunflecks is too short to initiate red light–triggered morphogenic reactions, because the activation of phytochrome is reversible by a rapid return to low ζ.

The scheme in Fig. 3.19 shows two phytochromes (A and B). Phytochrome B represents several phytochromes (B through E), which react in the same way and have been identified from several plant sources. While the interconversion of the physiologically inactive form (“Pr” for Pred) to the active form (“Pfr” for Pfar-red) by respective irradiation follows the same mechanism in all phytochromes, the inactivation mechanisms for Pfr differ. PhyAfr inhibits its own synthesis, and irradiation with red light as well as with blue light triggers its degradation. Synthesis and degradation play a secondary role in PhyB biochemistry as compared with photoconversion and slow reversion in the dark (Fig. 3.19). Note that photoconversion of phytochromes is associated with transport into (Pfr) and out of (Pr) the nucleus. The existence of a photoreceptor in two different states with distinct absorption maxima allows the monitoring of spectral quality (similar to colour vision in animals). Sunlight comprises both red and far-red light. Hence, the ratio of red to far-red light is crucial for the concentration of Pfr and the formation of a cellular signal.
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig19_HTML.png
Fig. 3.19

The red-light switch mediated by phytochrome. Phyfr is the active form of all phytochromes. (Modified from Weiler and Nover (2008))

Cryptochromes
Three cryptochromes have been found in A. thaliana, two acting primarily in the nucleus and one (Cry3) in mitochondria and chloroplasts. Cry3 is a photolyase with potentially some cryptochrome activity too (Liu et al. 2011). Cryptochrome 1 is the photoreceptor for high blue light intensity, while cryptochrome 2 is a sensitive blue light receptor that reacts at low fluence rates. The photoactive domains of the cryptochromes are homologous to those of the photolyases (N-terminal photolyase-homologous region (PHR)). However, because of their C-terminal extensions, cryptochromes do not exhibit photolyase activity (Fig. 3.20) but instead act as kinases. The photochemical mechanism(s) of cryptochrome light activation are not yet completely understood and the mechanisms of photoactivation and inactivation of both cryptochromes may differ to some extent (Li et al. 2011; Liu et al. 2011). Binding of the chromophores (5,10-methenyltetrahydrofolate (MTHF)) and FAD to the protein is non-covalent and the formation of an FAD radical anion (by electron transfer from MTHF) and a reduced FAD (FADH) radical (by subsequent proton transfer from the protein) appears obligatory in both cryptochromes. For activity, A. thaliana cryptochromes have to be phosphorylated. Upon additional (auto)phosphorylation after exposure to blue light, Cry2 (but not Cry1) is ubiquitinated and degraded. This mode of action resembles the situation in the phytochrome systems where PhyA is readily degraded, while PhyB is not (Liu et al. 2011).
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig20_HTML.png
Fig. 3.20

Cryptochromes are evolutionarily derived from ultraviolet (UV)-activated photolyases involved in DNA damage repair. Photolyases (type I: Escherichia coli; type II and 6–4 photolyase: Arabidopsis thaliana), their substrates and co-substrates. A blue-light-triggered intrinsic energy transfer from pteridine to reduced flavin adenine dinucleotide (FADH) leads to an energisation of the pyrimidine residues of the dimers (in a DNA strand) and to a rearrangement of the carbon bonds, resulting in the cleavage of the dimers and restoration of the original DNA helix. Structures of pterin (MTHF), deazaflavin (8-HDF) and the intermediate flavin radical (FADH). (Modified from Cashmore et al. (1999))

Phototropins

Phototropins (Phot 1 and Phot 2 in A. thaliana) are blue light receptors of the plasma membrane with two flavin mononucleotides (FMNs) as chromophores. In darkness, their C-terminal protein kinase domain is sterically inhibited. Absorption of blue light changes the protein conformation, leads to dissociation from the plasma membrane and unlocks the kinase activity. In the activated state, the FMN becomes covalently bound to a cysteine residue of the protein chain. This state of the protein is unstable and is readily reversed in the dark, whereupon the covalent bond breaks up. Along with cryptochromes and phytochromes, phototropins allow plants to respond to their light environment. For instance, they are important for the opening of stomata and mediate phototropic responses, e.g. the movements of chloroplasts (Fig. 3.4). Phot 1 is required for blue light–mediated transcript destabilisation of specific messenger RNAs (mRNAs) in the cell. For detailed chemical structures of the photoreceptors and mechanisms of gene regulation by light, see plant biochemistry textbooks (e.g. Buchanan et al. 2015).

3.2.5.2 Signal Transduction: The Signalling Centre COP1

Growth of plants in the dark, known as etiolation or skotomorphogenesis, requires suppression of photomorphogenesis, which is initiated by illumination. Activation of the above described photoreceptors - phytochromes and cryptochromes in particular - by irradiation triggers signal transduction cascades, which converge at a protein termed COP1 (CONSTITUTIVE PHOTOMORPHOGENIC 1). COP1 is a central switch which, because of its complex structure, can interact with several other proteins. In the dark, COP1—by its function as an E3 ubiquitin ligase—targets photomorphogenesis-promoting transcription factors for degradation by the ubiquitin-proteasome system, thus preventing photomorphogenesis. One of these proteins is the transcription factor ELONGATED HYPOCOTYL 5 (HY5) which, together with others, is responsible for the onset of photomorphogenesis (e.g. the inhibition of hypocotyl elongation growth). In the dark, it is degraded in the proteasome, giving rise to the well-known etiolated hypocotyls. Light-activated photoreceptors inactivate COP1 by mediating its export from the nucleus, thereby suspending its activity against transcription factors that are located in the nucleus. For a further understanding of the interaction of COP1 with the photoreceptors, it is necessary to introduce another protein, SPA1 (SUPPRESSOR OF PHY A 1), which forms a complex with COP1 to enable its association with the E3 ligase core complex; this finally ubiquitinates the proteins, targeting them for degradation (Fig. 3.21a). In the light, activated Cry1 and 2 interact with SPA1, dissociating it from COP1, which can now be exported from the nucleus (Fig. 3.21b). Further targets of the COP1 E3 ubiquitin ligase are the phytochromes—in particular PhyAfr and cryptochrome 2.
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig21_HTML.png
Fig. 3.21

Action of COP1 in the dark (a) and its inactivation by interaction with activated Cry1 in the light (b). The core protein complex for ubiquitination of proteins for degradation is composed of three proteins, which interact with the targeting dimer complex (COP1–SPA1)2 on the one hand and with the ubiquitinating enzyme E2 on the other. In the dark the transcription factor HY5 is ubiquitinated and degraded in the proteasome. Light-activated Cry1 dissociates the COP1–SPA1 interaction, releasing COP1 which, as a monomer, is exported from the nucleus. HY5 accumulates and initiates photomorphogenesis (Lau and Deng (2012))

While COP1 is a negative regulator of photomorphogenesis, it promotes UV-B-triggered reactions of plants, hence acting as a positive regulator (Fig. 3.22). To understand this difference, it is important to note that in skotomorphogenesis, COP1 acts as a homodimer, and also the COP1–SPA1 complex is dimeric. Dissociation of the COP1–SPA1 complex by sequestration of SPA1 through interaction with cryptochrome also monomerises the COP1 homodimer. In the cytosol, monomeric COP1 can bind to the monomeric activated UVR8 receptor and the heterodimer migrates back into the nucleus, where it activates HY5 and other transcription factors and effectors (Chap. 3, Sect. 3.4).
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig22_HTML.png
Fig. 3.22

Roles of the central regulator COP1 in signal transduction of skotomorphogenesis, photomorphogenesis a and ultraviolet (UV)-triggered responses b. In addition to photomorphogenesis, COP1 is involved in other developmental events (regulation of flowering, with the transcriptional regulator CO (CONSTANS)) and the circadian rhythm (by regulating the abundance of the circadian-associated protein GI (GIGANTEA)). (Modified from Lau and Deng (2012))

3.3 UV-B Radiation

3.3.1 Ranges of Ultraviolet Radiation and Biological Activity

The spectrum of solar ultraviolet light is continuous but is commonly divided into three wavelength bands: UV-C (200–290 nm), UV-B (290–320 nm) and UV-A (320–400 nm). UV-C is the most energetic of the three and is known as “germicidal UV” because of its potency against microorganisms (Yin et al. 2016). It is used to disinfect fresh fruit and vegetables to preserve their quality. However, since it is effectively absorbed by oxygen and ozone in the stratosphere, only a very small fraction reaches the Earth’s surface. UV-A, on the other hand, is not attenuated by atmospheric ozone, and this less damaging type of radiation plays an important role in plant photomorphogenesis. Although a sizeable amount of UV-B is absorbed by atmospheric ozone, its impact on life on our planet is considerable.

The development of the ozone hole—that is, the decrease in stratospheric ozone concentrations due to ozone decomposition by reaction with anthropogenic gases such as halogenated hydrocarbons or nitrogen oxides (IPCC/TEAP Special Report on Ozone and Climate 2005) is therefore observed with much concern. While the seasonally fluctuating ozone hole is particularly pronounced in the polar and subpolar regions, the UV radiation emitted by the sun crosses the atmosphere in regions of high geographical latitude at an angle and thus encounters a significantly “thicker” ozone layer than at the equator, where it takes the shortest path through the atmosphere. Therefore, in spite of the high latitudinal ozone hole, the UV radiation is high in the tropics and relatively low in the polar regions. The well-known altitudinal increase in UV radiation is caused by an attenuation of the tropospheric ozone layer, concomitant with a decrease in the intensity of haze.

UV-B radiation can damage cells and thus is dangerous to organisms (Table 3.3). In particular, plants—as sessile organisms—have developed mechanisms to cope with the ubiquitous and inescapable natural flux density of UV radiation through repair mechanisms for damaged cellular components (such as DNA). Protective measures are accumulation of UV-absorbing pigments in the epidermal layers, thick layers of hairs (Holmes and Keiller 2002; Manetas 2003) and cuticular waxes (Barnes et al. 1996). Generally, plants show adaptation to the UV-B load of their habitat. Plants growing along a latitudinal or an elevation gradient exhibit increased UV-B tolerance (Robberecht et al. 1980, Fig. 3.23). Problems caused by UV may arise for crop cultivars that are not well adapted to the natural UV-B stress occurring at their sites of cultivation, or for mobile organisms such as plankton. Phytoplankton, especially of the cold oceans, appear to be very sensitive to UV-B. In spite of the shallow depth to which UV light penetrates in a body of water (Fig. 3.24), significant reductions in phytoplankton biomass have repeatedly been observed as a consequence of the ozone hole. Likewise, decreases in yields have been reported for UV-sensitive crop cultivars (e.g. of maize and soybeans (Rius et al. 2016)). However, although statistically significant, these reductions were relatively small (usually <10%).
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig23_HTML.png
Fig. 3.23

Efficiency of ultraviolet (UV)-B screening (percent transmittance of impinging UV-B radiation) through the epidermis of several plant life forms from various provenances. (Modified from Munk (2009), Körner (1999) and Day et al. (1993))

/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig24_HTML.png
Fig. 3.24

Decrease in ultraviolet (UV) radiation with depth of water a in the northern Adriatic and b in the humin-rich Lake Neusiedler (Austria). Both measurements were conducted on a cloud-free August day. The intensity of the UV radiation at a depth of 5 cm in Lake Neusiedler corresponded approximately to a depth of 5 m in the Adriatic. (Modified from Herndl (1996))

Table 3.3

Physiological effects of increased ultraviolet (UV)-B radiation. (Modified from Jansen et al. (1998))

Damage

DNA damage

Dimerisation of thymine; strand breaks

Biomembranes

Lipid peroxidation

Photosynthetic apparatus

Inactivation of photosystem II; acceleration of D1 turnover; damage to thylakoid membranes; bleaching of pigments; decrease in the activity of photosynthetic enzymes (particularly RubisCO); inactivation of photosynthetic genes

Phytohormones

Photo-oxidation of auxin

UV avoidance

Activation of secondary metabolism

Activation of expression of the key genes of phenylpropanoid metabolism and accumulation of flavonoids; accumulation of alkaloids, waxes and polyamines

Production of radical scavengers

Increased capacity of the anti-oxidative system (ascorbate peroxidase, superoxide dismutase, glutathione reductase and others, e.g. mycosporines)

RubisCO ribulose-1,5-bisphosphate carboxylase/oxygenase

Morphological–anatomical symptoms indicating a still tolerable UV-B stress are swollen and shortened internodes, reduced leaf expansion, curling up of leaf edges and enhanced branching of the shoot through promotion of lateral buds. Leaves tend to show succulence with a particularly thick, usually pigmented epidermis and low density of stomata. Tolerated UV stress often results in accumulation of vitamin C and soluble sugars in fruits as radical scavengers, thereby also increasing the quality of food. However, the transition from UV-B-triggered “normal morphogenetic reactions” to those that have to be considered as stress responses is not clearly apparent, inasmuch as other natural stresses such as drought, nutrient deficiency or low temperature can interact with UV-B effects at the molecular level. Nevertheless, differentiation between normal development of UV-B tolerance and stress responses is clear when above-ambient levels of UV-B cause damage to DNA, proteins and membrane lipids, and inhibit protein synthesis and photosynthetic reactions. However, since UV-B also generates ROS (mainly the superoxide radical O2 ) through its impact on photosynthesis, respiration and on enzymes such as peroxidases and oxidases, the origin of UV-B triggered damage is not always obvious.

Investigation of UV-B effects on plants is not trivial, because of the omnipresence of this radiation in nature and, to a lesser extent, in artificial light sources. Many experiments have therefore applied pulses of UV-B overdoses or longer than natural exposure periods. Such treatment can easily overstretch the tolerance—for example, the capacity for repair—of the plants, causing unnatural reactions and even necroses and death. Also, different reactions have been observed, depending on whether the same total overdose of UV-B was applied either in high intensity pulses or as a slightly elevated constant flux over a long period. With the UV stress applied in short pulses, damage and repair are the dominant effects, while continuous but low-intensity stress leads to acclimation and damage avoidance.

3.3.2 Ultraviolet-B Damage and Repair Mechanisms

DNA has a broad absorption peak between 235 and 315 nm and thus is photoactivated by UV. Since the effects of UV-B radiation on DNA affect all kinds of organisms, they have been particularly well studied. Several types of UV-triggered damage are known: strand breaks and cross-linking, as well as modifications of pyrimidine bases. In plants, dimerisation of thymine—resulting in cyclobutane pyrimidine dimers (CPDs) and, to a smaller extent, in pyrimidinone dimers (known as 6–4 photoproducts)—is a common reaction to irradiation with UV-B. Less packaged DNA of mitochondria and probably also of chloroplasts is particularly sensitive. UV irradiation of yeast cells caused a 10% loss of nuclear DNA, but at the same time a 50–60% loss of mitochondrial DNA.

There are several mechanisms to repair such damage (for details, see biochemistry textbooks, e.g. Berg et al. (2015)), of which two are mentioned here: the DNA photolyase reaction, which requires blue light/UV-A; and light-independent reactions such as base excision and recombination. Repair usually takes only a few hours. Interestingly, UV-activated photolyases represent the evolutionary origin of the blue light receptor cryptochrome (Fig. 3.20). Strong UV stress leads directly to irreparable chromosome breakage and deletions resulting in the death of the organism. This is the basis for sterilisation of rooms and instruments with intense UV light.

Irrespective of the repair mechanisms, DNA damage signalling involves a UV-damaged DNA-binding protein complex (UV–DDB complex), which recognises UV-induced DNA damage and recruits proteins of the nucleotide excision repair pathway. It also induces several other UV-B responses, such as the formation of protective pigments.

In addition to UV damage to DNA, photo-oxidation of UV-absorbing pigments is known: yellowing and complete bleaching of leaves of indoor plants after abrupt transfer to the open air are frequent phenomena. Here the protective effect of the chloroplast pigments (energy dissipation) on their protein environment can be seen. Photodestruction of the thylakoid pigments leads to a significant decline in the amount of ribulose-1,5-bisphosphate carboxylase/oxygenase (RubisCO).

3.3.3 Avoidance of Ultraviolet-B-Induced Stress

Plants synthesise a range of compounds in order to avoid damage by UV-B radiation. The rate of accumulation in different plant types correlates well with the average UV exposure associated with the respective habitat (Fig. 3.23). Phenylpropanes, as aromatic compounds, exhibit strong absorbance of shortwave radiation and are thus able to function as effective UV filters or sunscreen pigments. Many compounds in this family of secondary plant metabolites absorb light only in the UV range and cannot be recognised as pigments by the human eye. Others, such as the anthocyanins, absorb light also in the visible range and therefore appear as pigments (Figs. 3.25 and 3.26).
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig25_HTML.png
Fig. 3.25

Anthocyanins in flushing leaves of the tropical Sapotacea Inhambanella henriquesii in the coastal forests of Kenya. The anthocyanin is located in the vacuoles of the leaf epidermis and protects the leaves, which are not yet fully green, from radiation damage. The close-up (right) shows the mixture of colours between the red anthocyanins and the chlorophyll. (Photos: E. Beck)

/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig26_HTML.png
Fig. 3.26

After glycosylation, phenylpropanes are accumulated in the vacuoles of the epidermal cells. a Chemical structure of the anthocyanidin malvidin (at acidic pH, arrows indicate the positions for glycosylation) and b the absorption spectrum of its 3-O-glucoside (which is nearly identical to that of the aglycone malvidin) in comparison with the absorption spectra of chlorophylls a and b

Glycosylation renders phenylpropanes water soluble so they can be sequestered in the vacuoles primarily of the epidermis. In certain purple varieties such as copper beech (Fagus sylvatica), the vacuoles of the mesophyll cells contain such pigments too. They absorb in the UV spectrum (Fig. 3.26) and in the green region of the visible spectrum and thus do not interfere with the absorption of light by the chlorophylls. Protective colouration of the leaves by anthocyanins is frequently observed in unfolding young leaves (called “juvenile anthocyanin”) and again during senescence. In young leaves the additional pigments protect the developing photosynthetic machinery until the energy dissipation mechanisms are fully functional. The “autumn anthocyanin” is well known from the Indian-summer aspect of deciduous trees. Its ecological function, however, is not well understood. Most likely it protects the controlled degradation of leaf constituents—in particular, the chlorophylls—and the export of degradation products to the overwintering parts of the plant. The key enzyme in the flavonoid metabolism is chalcone synthase, whose formation is strongly induced by UV-B as well as by UV-A/blue light.

Other examples of compounds—out of the great variety of aromatic secondary metabolites—that accumulate upon UV-B exposure include 3″,6″-DCA and 3″,6″-DCI (=3″,6″-di-para-coumaroyl-astragalin and 3″,6″-di-para-coumaroyl-isoquercitrin) in pine needles and 2″,2″-di-para-coumaroylkaemperol-3α-d-arabinoside in beech leaves (Figs. 3.27 and 3.28). The non-conjugated phenylpropanes (e.g. p-coumaric acid and its derivatives) are primarily localised in the epidermal cell wall, while the conjugates are mainly sequestered in the vacuoles of the same cells. It has been calculated that the accumulation of aromatic compounds in the outer epidermis cell wall of pine needles allows only about 4% of the incoming UV-B to pass through. Diacylated flavonoids in the vacuole are able to filter out the residual UV-B, leaving the mesophyll completely unaffected. This explains the exceptional UV protection that needles of coniferous trees are known for (Fig. 3.23).
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig27_HTML.png
Fig. 3.27

Histochemical evidence for ultraviolet (UV)-absorbing substances in the epidermis of pine needles. ac Fluorescence (excited by light of 450–490 nm wavelength, determined at >520 nm): a control; bc after treatment with a reagent that intensifies flavonoid fluorescence. df Observation with a confocal laser microscope (which shows the cells and their content more clearly): (e) control; df after staining, as above. abef Transverse section of a needle; cd stomata. (Modified from Schnitzler et al. (1996))

/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig28_HTML.png
Fig. 3.28

Examples of diacylated flavonoids. 3″,6″-Di-p-coumaroyl-astragalin (DCA, R=H) and 3″,6″-Di-p-coumaroyl-isoquercitrin (DCI, R=OH)

The association between concentrations of UV screens and UV exposure in the natural habitat is apparent when comparing, for instance, alpine herbs with herbs from the lowlands (Filella and Peñuelas 1999; Fig. 3.23). Moreover, because synthesis of UV-B screening phenols and phenylpropanes is stimulated by UV-B radiation, variations in their concentrations can be expected also when leaves of the same species but from locations with contrasting UV-B radiation intensities are compared. One example of such a local adaptation is mountain avens (Dryas octopetala) from the Arctic, from southern Norway and from the French Alps, where the average UV-B radiation is three times higher than in the Arctic. UV-B transmittance in the epidermis of Dryas from the French Alps was only 2.5% versus 5% in plants from Norway and 7% in Arctic plants (Nybakken et al. 2004). It was also shown that depending on the weather conditions during the year, the screening capacities of the leaf epidermis may change, indicating a dynamic acclimation response (Barnes et al. 1996).

3.3.4 Ultraviolet-B Perception and Signalling

Many transcriptional responses to increased doses of UV-B are non-specific and shared with, for instance, defence responses against pathogen attack (Stratmann 2003). The respective signalling has been associated with UV-B as a severe stressor of plant life, is triggered by DNA damage and ROS production, and involves typical plant stress hormones (Fig. 3.29). Responses result in repair and in protection by the same classes of chemical compounds, which can act either as sunscreen pigments or as components for defence (phytoalexins). In particular, induction or reinforcement of radical scavenging systems such as glutathione reductase, ascorbate peroxidase and glutathione peroxidase are very important for avoidance and alleviation of UV-triggered stress. Overall, responses to UV-B stress are dependent on signalling pathways triggered by non-specific secondary stress (Chap. 2, Sect. 2.​2), while photomorphogenic acclimation responses are more UV-B specific and are triggered by low levels of UV. It is important to note that both contribute to survival. The specific responses can—depending on the wavelength, intensity or duration—also represent acclimation and avoidance reactions. During the past decade, much progress has been made in the elucidation of signal transduction chains resulting in UV-B tolerance of plants.
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig29_HTML.png
Fig. 3.29

Ultraviolet (UV)-B signal transduction pathways. UV-B induces UV-B-specific and non-specific signal production and transduction. The UV-B wavelength, intensity and duration of exposition trigger the induction of specific sets of target genes and downstream responses that result in repair of UV-triggered damage and adaptation to UV-B. JA jasmonic acid, SA salicylic acid. (Modified after Jenkins (2009))

Photomorphogenic signalling requires perception of radiation. While the photoreceptors for visible light contain chromophores, the UV-B photoreceptor lacks such a component. It absorbs shortwave radiation by a series of aromatic amino acid residues—in particular, 14 tryptophan residues (Trp, W). The UV-B receptor was originally described as a regulatory protein in UV-triggered signal transduction in A. thaliana (Kliebenstein et al. 2002) and was termed UV RESISTANCE LOCUS 8 (UVR8). In 2011, Rizzini et al. identified this protein as the long-sought-after UV-B photoreceptor which, by a special arrangement of tryptophan residues and positively charged arginine residues, can convert UV-B radiation into a chemical signal. Other authors contributed the detailed photochemical and biochemical mechanisms of that process (Christie et al. 2012; Wu et al. 2012). In all investigated species, from algae to higher plants, this type of UV-B photoreceptor has been found (Rizzini et al. 2011).

In the energetic ground state (e.g. in the dark), UVR8 is a doughnut-shaped homodimer whose monomers are linked by a network of salt bridges and aromatic side-chain interactions (Gardner and Correa 2012) (Fig. 3.30). Each monomer consists of 440 amino acid residues with 14 Trp residues, which are clustered at the top surface where the dimer forms. Three of them form a triad, which interacts with another tryptophan of the counterpart analogue (Fig. 3.30). Excitation of the tryptophan residues by UV-B results in the dissociation of the salt bridges and releases the monomers. The active UVR8 monomer then binds to the multifunctional COP1 protein, which is a central regulator in UV-B and visible light signalling (Oravecz et al. 2006) (Fig. 3.22). At this point the question of “normal photomorphogenesis” versus the UV stress response arises again but cannot be conclusively answered. It is, however, clear that UV-B-specific responses—whether they are contributing to the normal development of a plant or are excited by an unusual dose of UV radiation—are mediated by the UVR8COP1 signalling pathway (Heijde and Ulm 2012). The central protein of this pathway is COP1, best known as a negative regulator or repressor of photomorphogenesis (Fig. 3.22). However, in UV-B signalling, COP1 is a positive regulator. The interaction complex of monomeric UVR8 with COP1 migrates from the cytosol into the nucleus, where it activates HY5 gene expression. The transcription factor HY5 triggers the expression of a great variety of UV-B-responsive genes. Among these are genes that encode proteins required for UV-B tolerance or protection, such as photolyases for DNA repair, and enzymes of the phenylpropanoid pathway, such as chalcone synthase, by which phenolic UV-B scavengers are produced.
/epubstore/S/E-D-Schulze/Plant-Ecology/OEBPS/images/72100_2_En_3_Chapter/72100_2_En_3_Fig30_HTML.png
Fig. 3.30

Plant ultraviolet (UV)-B sensing by the UVR8 photoreceptor. UVR8 exists as a stable dimer in the dark through salt bridges and interactions of tryptophan residues forming a Trp-pyramid (3 + 1 trp residues •). Excitation of the tryptophan residues by UV-B results in the donation of π electrons to the arginine residues of the salt bridges, leading to charge neutralisation and dissociation of the dimer. Monomeric UVR8 binds to COP1 which, as a complex, can migrate into the nucleus for expression of the HY5 transcription factor and other genes that lead to the formation of, for example, UV screens. Among these genes is that for RUP proteins (REPRESSOR OF UV-B PHOTOMORPHOGENESIS), which can replace COP1 and finally, by dissociation from UVR8, facilitate dimerisation of (inactive) UVR8. (Modified from Gardner and Correa (2012); for details, see Jenkins (2014))

3.3.5 Crosstalk Between Ultraviolet-B and Visible Light Responses

COP1 interacts with all known photoreceptors for visible and UV-B light in specific ways, thereby enabling crosstalk between the photoreceptors. The multifunctionality of the COP1 protein is based on its complex structure, which consists of three major domains: a RING finger, a coiled coil domain and a so-called WD40 (TrpAsp) repeat domain. COP1 can dimerise through the coiled coil domain. Furthermore, the photoreceptors for red/far-red (phytochromes), blue (constitutive cryptochrome effect) and UV-B (UVR8) interact with specific sites of the WD40 domain. By the versatility of the interactions of photoreceptors with the same central regulatory protein, COP1 specificity of signalling can be achieved. While, for example, the ubiqitination and degradation of the photolabile receptors Cry2 and PhyA is mediated by COP1 (Fig. 3.22), it does not impair the stability of PhyB, Cry1 and UVR8. Further regulation of COP1 activity and specificity is achieved by interaction of the coiled coil domain with proteins of the SUPPRESSOR OF PHYTOCHROME A family (SPA1–SPA4; Fig. 3.21), which are required for COP1 function in the dark and visible light but not for the UV-B response. On the other side, REPRESSOR OF UV-B PHOTOMORPHOGENESIS 1 and 2 (RUP1 and RUP2, Fig. 3.30) have been identified as negative feedback regulators of UVR8 signalling (Gruber et al. 2010).

3.4 Summary

  • For plants, light has a dual function as an energy source and as a signal crucial for plant development and adaptation to the environment. The two basic functions are mediated by light-absorbing molecules, which differ in spectral sensitivity, biological activity and subcellular localisation: the light-harvesting systems for photosynthesis are in the chloroplasts, and the light sensors regulating developmental processes are predominantly cytosolic and/or in the nucleus.

  • Light as a stressor: An unfavourable light environment may result from low and high light intensity, as well as from rapid changes between both radiation intensities. Plants can acclimate to their light environment at the level of life strategies, by positioning, morphology and structure of leaves and chloroplasts, as well as by physiological, biochemical and biophysical adjustments.

  • While the interpretation of morphological, structural and physiological acclimations to a stressful light environment is obvious, the mechanisms by which the permanent or dynamic acclimations are achieved are understood in less detail, except in a few cases, such as light harvesting and energy dissipation, or the shade avoidance response of juvenile plants as an example of development guided by the light conditions.

  • Only a small portion of the visible radiation absorbed by photosynthetic pigments can be used for photosynthetic CO2 assimilation. Under otherwise adverse environmental conditions such as a shortage of water, this portion may be as low as 1% or 2%. Overexcitation results in damage to the photosystems—mainly photosystem II—either directly or by the formation of radicals and reactive oxygen species (ROS). Several mechanisms are known for avoidance of overexcitation or for dissipation of excess light energy.

  • The most important mechanism for energy dissipation is the so-called non-photochemical quenching (NPQ or qE), which takes place in the antenna system. Conformational change of the proton-sensing protein PsbS of the thylakoid membrane switches between the light harvesting and energy dissipation modes of photosystem II. The xanthophyll cycle participates in the latter reaction. Energy-dependent dissipation of the peripheral antennae from photosystem II and partial association with photosystem I (the state I–state II transition) provide for a balanced excitation of both photosystems in the case of high irradiation.

  • Acclimative changes between the harvesting and the dissipative modes of the photosystem are necessary for leaves of a tree crown during the course of a day. Acclimated to a moderate light intensity, they are transitorily stressed by direct radiation, termed sunflecks. The photosynthetic gain from such a changing light environment is lower than from continuous illumination.

  • Continuous light, as applied in greenhouses, is injurious for many plant species, as it is not natural. Even in the polar summer, light intensities oscillate considerably in a circadian rhythm.

  • Perception of light by a range of receptors and subsequent translation of the respective signals control many aspects of plant development and are crucial for acclimatisation to environmental stressors in diverse habitats. Phytochromes are receptors for red and far-red light, and cryptochromes and phototropins are receptors for blue light. The photoreceptors for red and blue light principally consist of a protein and a chromophore. Because of the UV absorption by aromatic amino acid residues, the photoreceptor for UV-B light does not require a chromophore.

  • Two classes of phytochromes are known—PhyA and PhyB—the latter comprising several species, which are differentiated by their proteins (PhyB–E). Physiologically inactive phytochromes (Phyred) of both classes absorb red light by an open-chain tetrapyrrole chromophore and change their structure to the physiologically active form with an absorption peak in the infrared (Phyfar-red). Phyfar-red migrates into the nucleus for triggering activity, while Phyred is localised in the cytosol. Both classes of phytochromes are differentiated by their mode of inactivation: PhyAfar-red inhibits its own synthesis, and irradiation with red as well as with blue light triggers its degradation. PhyBfar-red slowly reconverts into PhyBred (in the dark) or is degraded.

  • Cryptochrome 1 is the photoreceptor for high blue light intensity, while cryptochrome 2 is a sensitive receptor that reacts at low intensities. The photoactive domains of cryptochromes apparently evolved from photolyases but, because of their C-terminal extensions, they act as kinases and do not exhibit photolyase activity. Their chromophores are pterins and flavins. For activity, cryptochromes must become phosphorylated. Further (auto)phosphorylation triggers degradation in the proteasome and thus stops cryptochrome activity.

  • Phototropins are blue light receptors of the plasma membrane with two flavin mononucleotides (FMNs) as chromophores. Absorption of blue light changes the protein conformation, leads to dissociation from the plasma membrane and unlocks a kinase activity, which activates regulatory proteins. Phototropins are important for many blue-light-dependent reactions of the plant—for example, opening of stomata (activation of the proton pump) and phototropic responses.

  • Activation of the photoreceptors triggers signal transduction cascades, which merge at the central protein COP1 (CONSTITUTIVE PHOTOMORPHOGENIC 1) which, as a homodimer, can interact with several other proteins. In the dark, COP1 exhibits activity as E3 ubiquitin ligase, targeting photomorphogenesis-promoting transcription factors for degradation in the proteasome and thus preventing photomorphogenesis. Light-activated photoreceptors inactivate COP1 by mediating its monomerisation and export from the nucleus, thereby suspending its activity against transcription factors in the nucleus. Further targets of the COP1 E3 ubiquitin ligase are also the phytochromes—in particular, PhyAfar-red and cryptochrome 2.

  • Under irradiation with UV-B, monomeric COP1 associates with the activated UV-B receptor. The heterodimer is transported back into the nucleus, where it triggers expression of genes for photomorphogenesis and UV protection.

  • The spectrum of ultraviolet radiation comprises three wavelength ranges according to the energy level: UV-C (100–280 nm), UV-B (260–315 nm) and UV-A (315–400 nm). The shorter the wavelength is, the stronger is its destructive effect, but the more is also absorbed by ozone in the stratosphere and the troposphere. UV-C is almost completely absorbed when passing through the atmosphere; UV-A passes through the atmosphere but hardly damages organisms. The biologically most effective UV light is UV-B. Due to the fluctuating attenuation by stratospheric ozone (the “ozone hole”), the intensity of UV-B reaching the Earth’s surface can increase temporarily.

  • Strong UV-B radiation damages plants, particularly DNA (breakage of strands and deletions) in the nucleus, the chloroplasts and mitochondria. It also damages the photosynthetic apparatus (bleaching of photosynthetic pigments, destruction of photosynthetic proteins). Organisms have very efficient repair systems, particularly for damaged DNA (photolyases and nucleotide excision).

  • Plants have a high potential to develop protection against UV-B stress. By accumulating UV-B screen compounds, plant can adapt and/or acclimate to UV-B impact. The degree of UV protection correlates well with the extent of UV exposure typical of the respective habitats.

  • UV-B screens consist, above all, of strongly UV-absorbing phenylpropanoid pigments in the walls and vacuoles of epidermal cells. Common radical scavengers, such as the ascorbate peroxidase system or glutathione, are also boosted because of the ROS production elicited by UV stress.

  • UV responses involve non-specific and specific signalling pathways. The latter are triggered by damaged DNA or ROS. Specific responses depend on UV-B-specific receptors such as Arabidopsis thaliana UVR8. Morphogenetic UV-B signalling can interact with signals from phytochrome and cryptochrome. Crosstalk is integrated by the multifunctional COP1 protein, which controls protein degradation.