Plasticity in light reactions of photosynthesis for energy production and photoprotection (2024)

Article Navigation

Volume 56 Issue 411 January 2005

Article Contents

  • Abstract

  • Photosynthesis and its down-regulation

  • The need for modulation of down-regulatory sensitivity (qE-modulation)

  • The need to balance ATP/NADPH ratios

  • Type I: Flexibility mechanisms affecting the ATP/NADPH ratio

  • Type II: Flexibility without altering ATP/NADPH output ratio

  • An integrated model for photosynthetic flexibility

  • References

  • < Previous
  • Next >

Journal Article

,

Jeffrey A. Cruz

*To whom correspondence should be addressed. Fax: +1 509 335 7643. E-mail: dkramer@wsu.edu

Search for other works by this author on:

Oxford Academic

,

Thomas J. Avenson

Search for other works by this author on:

Oxford Academic

,

Atsuko Kanazawa

Search for other works by this author on:

Oxford Academic

,

Kenji Takizawa

Search for other works by this author on:

Oxford Academic

,

Gerald E. Edwards

Search for other works by this author on:

Oxford Academic

David M. Kramer

Search for other works by this author on:

Oxford Academic

Journal of Experimental Botany, Volume 56, Issue 411, January 2005, Pages 395–406, https://doi.org/10.1093/jxb/eri022

Published:

08 November 2004

Article history

Received:

03 June 2004

Accepted:

31 August 2004

Published:

08 November 2004

  • PDF
  • Split View
  • Views
    • Article contents
    • Figures & tables
    • Video
    • Audio
    • Supplementary Data
  • Cite

    Cite

    Jeffrey A. Cruz, Thomas J. Avenson, Atsuko Kanazawa, Kenji Takizawa, Gerald E. Edwards, David M. Kramer, Plasticity in light reactions of photosynthesis for energy production and photoprotection, Journal of Experimental Botany, Volume 56, Issue 411, January 2005, Pages 395–406, https://doi.org/10.1093/jxb/eri022

    Close

Search

Close

Search

Advanced Search

Search Menu

Abstract

Plant photosynthesis channels some of the most highly reactive intermediates in biology, in a way that captures a large fraction of their energy to power the plant. A viable photosynthetic apparatus must not only be efficient and robust machinery, but also well integrated into the plant's biochemical and physiological networks. This requires flexibility in its responses to the dramatically changing environmental conditions and biochemical demands. First, the output of the energy-storing light reactions must match the demands of plant metabolism. Second, regulation of the antenna must be flexible to allow responses to diverse challenges that could result in excess light capture and subsequent photoinhibition. Evidence is presented for the interplay of two types of mechanistic flexibility, one that modulates the relative sensitivity of antenna down-regulation to electron flow, and the other, which primarily modulates the output ratio of ATP/NADPH, but also contributes to down-regulation.

ATP synthase, cyclic electron transfer, proton motive force, qE quenching, water–water cycle

Photosynthesis and its down-regulation

Light is captured by a set of light-harvesting complexes (LHCs) that funnel light energy into photochemical reaction centres, photosystem (PS) I and PSII (Fig. 1) (see review by Ort and Yocum, 1996). Special subsets of chlorophyll molecules in these photosystems are excited by light energy, allowing electrons on them to be transferred through a series of redox carriers called the electron transfer chain (ETC), beginning from the oxygen evolving complex (OEC) of PSII (which oxidizes H2O and releases O2 and protons) (Diner and Babco*ck, 1996), through the plastoquinone (PQ) pool, the cytochrome (cyt) b6f complex (Sacksteder et al., 2000) and plastocyanin (PC), and finally through PSI (Malkin, 1996). Electrons from PSI are transferred to ferredoxin (Fd), which, in turn, reduces NADP+ to NADPH via ferredoxin:NADP+ oxidoreductase (FNR) (Knaff, 1996). This linear electron flux (LEF) to NADP+ is coupled to proton release at the OEC, and ‘shuttling’ of protons across the thylakoid membrane by the PQ pool and the Q-cycle at the cyt b6f complex, which establishes an electrochemical potential of protons, or proton motive force (pmf) that drives the synthesis of ATP by chemiosmotic coupling through the chloroplast ATP synthase (McCarty, 1996; Mitchell, 1966).

Fig. 1.

Plasticity in light reactions of photosynthesis for energy production and photoprotection (3)

Open in new tabDownload slide

Primary routes of proton/electron flux and mechanisms of Type I and II flexibility. (A) Energy storage begins with the absorption of light energy (lightning bolts) by light-harvesting complexes (LHC) associated with photosystem (PS) II and I, respectively. Depicted is the linear electron flux (LEF, red arrows) of electrons derived from the oxidation of H2O at the oxygen evolving complex (OEC) through PSII reducing sequentially plastoquinone (PQ) to a quinol (PQH2). Bifurcated oxidation of PQH2 occurs at the cytochrome b6f complex (b6f ) where half of the electrons are linearly transferred to the NADP+/NADPH couple via plastocyanin (PC), PSI, ferredoxin (Fd), and ferredoxin-NADP+ oxidoreductase (FNR), and the other half of the electrons will return to the PQH2 pool. Proton flux (blue arrows) originates from H2O splitting at the OEC and the cyclic reduction and oxidation of PQ/PQH2, establishing an electrochemical gradient of protons across the thylakoid membrane (pmf), comprised of pH (ΔpH) and electric field (Δψ) components. Total pmf drives ATP synthesis from ADP and Pi as protons move down their electrochemical gradient through the CF1-CFO ATP synthase. Energy dissipation by qE (purple arrow) is pH-dependent due to the pH-dependent activity of violaxanthin de-epoxidase (VDE), which sequentially reduces violaxanthin (V) to zeaxanthin (Z), and protonation of PsbS. Type II mechanisms (highlighted in red) involve variability in: (i) the response of the antenna to lumen pH, (ii) the conductive properties of the ATP synthase, and (iii) the relative partitioning of pmf into Δψ and ΔpH. Type I mechanisms (B) involve alternate routes of electron transfer at the reducing side of PSI, including the water–water cycle (WWC) and cyclic electron flow around PSI (CEF1). The WWC uses the same electron transfer pathways as normal LEF except at the reducing side of PSI it reduces O2 to

\(\mathrm{O}_{2}^{{-}}\)

which is subsequently detoxified to H2O. As depicted, four carrier pathways have been proposed for the cycling of electrons from PSI back to the PQ pool (CEF1): (1) a ferredoxin-PQ oxidoreductase (FQR), (2) a NADPH-PQ oxidoreductase (NDH), (3) oxidation of Fd by a FNR/b6f super complex, and (4) oxidation of, for example, Fd by a newly discovered haem associated with the stromal side of the b6f complex.

Following the absorption of photons by chlorophyll, the transfer of excitons to reaction centre chlorophyll and the initiation of electron transfer must be well regulated to prevent ‘over-excitation’ of the photosystems (i.e. more excitation than can be processed by the reaction centres), which favours the formation of highly reactive species and photoinhibition of the photosynthetic machinery (Anderson and Barber, 1996; Kramer and Crofts, 1996). In general, overexcitation of PSII is prevented largely by antenna down-regulation, which dissipates excess excitation energy as heat. This involves a series of processes, which are collectively termed non-photochemical exciton quenching (NPQ) and typically measured by the quenching of chlorophyll a fluorescence (reviewed by Maxwell and Johnson, 2000). Under most physiological conditions, the major form of NPQ is termed qE, for the ‘quenching’ of light energy in the antenna that is dependent on the ‘energization’ of the thylakoid membrane (reviewed by Horton et al., 1996; Müller et al., 2001; Owens, 1996; Yamamoto and Bassi, 1996). Activation of qE involves at least two processes (Fig. 1): (i) the conversion of the xanthophyll carotenoid violaxanthin to antheraxanthin and zeaxanthin, catalysed by violaxanthin deepoxidase (VDE) (Eskling et al., 2001); and (ii) protonation of amino acid side-chains on an antenna-associated, chlorophyll binding protein, PsbS (Li et al., 2004). Both of these processes are activated by acidification of the lumen by the ΔpH component of pmf. In this analysis, the proton gradient is considered to equilibrate across the entire continuous lumenal space, i.e. it is not necessary to invoke proton domains to explain these data. Thus, pmf not only drives the synthesis of ATP, but is also a key signal for feedback regulation of the light reactions.

The need for modulation of down-regulatory sensitivity (qE-modulation)

qE sensitivity is defined as the responsiveness of qE to LEF, because both parameters are readily and frequently measured using chlorophyll fluorescence measurements. Alternatively, under most conditions NPQ may be substituted for qE, since qE makes up a significant fraction of NPQ. If the light reactions behaved in a static fashion, qE sensitivity would be constant, i.e. qE would be a continuous function of LEF. However, such rigidity in down-regulation of the photosynthetic apparatus would leave it prone to catastrophic failure (Asada, 1996; Heber and Walker, 1992; Kanazawa and Kramer, 2002). For example, if photosynthesis became limited by the lack of PSI electron acceptors, as might be expected under conditions of metabolic stress, LEF and its proton pumping will be attenuated. A static model would predict a decrease in qE, precisely under the conditions where photoprotection is needed most to prevent the build-up of reduced intermediates, which could lead to ‘acceptor side’ photoinhibition (Anderson et al., 1997). Clearly, a more flexible model must be invoked to account for the response of antenna regulation to the fluctuating physiological status of the plant (Horton, 1989; Horton et al., 1999).

Indeed, such flexibility has been demonstrated in C3 plants (Avenson et al., 2004; Kanazawa et al., 2001; Kanazawa and Kramer, 2002). Rather than a continuous relationship, as the static model would predict, a series of distinct curves was observed, with qE becoming increasingly more sensitive to LEF as [CO2] was lowered (Kanazawa and Kramer, 2002). Physiologically, this is desirable because the availability of PSI electron acceptors, and thus overall LEF, is expected to decrease with decreasing CO2; to maintain reasonable levels of photoprotection, qE should become more sensitive to LEF.

The need to balance ATP/NADPH ratios

With LEF to NADP+, ATP synthesis and NADPH production are coupled, and within a static model the output ratio of ATP to NADPH would be fixed. However, this would work only in a system where consumption of ATP and NADPH occurs at the same fixed ratio; that is, their relative consumption by chloroplast metabolism (including fixation of carbon, nitrogen, phosphorus, and sulphur) and other plastid maintenance processes continuously matches output by energized thylakoids. Yet each individual process imposes a different demand for ATP/NADPH. Again, this leaves a static model susceptible to failure in cases where differential flux is required to respond to the changing demands on the chloroplast. If shortage of a single metabolite decreases relative metabolic flux through the pathway that fixes it, then the demand for ATP versus NADPH may change. Also, the resulting mismatch between production and consumption ratios would create ‘back pressure’ on the light reactions from excess product (ATP or NADPH) or lack of substrate (ADP+Pi, NADP+), sensitizing the photosynthetic apparatus to photoinhibition. Therefore, contrary to a static model, a certain measure of flexibility in the LEF output ratio is expected in order to compensate for changes in demand.

The need for balancing mechanisms is further exemplified by potential mismatch between the LEF-dependent ouput and the demand of CO2 fixation. If one considers only LEF, the ATP/NADPH ratio is defined by the proton coupling stoichiometries for the ETC (H+/e) and that for ATP synthesis (H+/ATP, termed n) (Allen, 2002; Kramer et al., 2003). There is strong evidence that H+/e for LEF remains at 3 under physiological conditions (Sacksteder et al., 2000). New information about the structure and mechanism of the ATP synthase implies that n is likely to be 4.67 (reviewed by Allen, 2002; Kramer et al., 2003). With these stoichiometries, ATP/NADPH should be 1.3, which as discussed later, would provide insufficient ATP to support CO2 fixation in C3 plants. Without flexible responses, even larger supply–demand mismatch would occur in species using modified CO2 fixation strategies, for example, in plants with some types of C4 photosynthesis.

As discussed in (Kanazawa and Kramer, 2002) and extended here, there are several models that could, together or separately, account for qE modulation (Fig. 1), some of which will also affect the output ratio of ATP/NADPH, and these were termed Type I flexibility mechanisms. Other mechanisms will have no effect on the ATP/NADPH output ratio, and these were termed Type II flexibility mechanisms. This distinction is critical for understanding the relative roles of these processes.

Type I: Flexibility mechanisms affecting the ATP/NADPH ratio

In accordance with the general model for electron and proton transfer, any process increasing the rate of proton translocation into the lumen will tend to activate qE by increasing pmf. If such processes supplement proton flux supplied by LEF, they will increase qE sensitivity as it has been defined here. They will also tend to increase the ATP/NADPH ratio, because the resulting increase in proton flux will drive more ATP synthesis (Fig. 2), without a net increase in the reduction of NADP+. There have been several proposals for this type of mechanism.

Fig. 2.

Plasticity in light reactions of photosynthesis for energy production and photoprotection (4)

Open in new tabDownload slide

Relationships between energy-transduction and qE sensitivity. As determined by its sensitive components, PsbS and VDE, qE (and thus NPQ) will be a function of lumenal pH. As pH drops from ∼6.5 to ∼5.8, qE will continuously increase to saturation. If the steady-state pH of the stroma is constant, then qE will be a function of ΔpH. Therefore, factors affecting the extent to which ΔpH forms will influence qE induction. Depicted are simplified schematics of chloroplastic energy transduction with proton and electron fluxes indicated in blue and red, respectively. The table indicates relative changes in ATP output, NADPH output, pmf, and ΔpH (NC indicates no change). The pmf (and by extension ΔpH) will depend, in part, on the steady-state rate of proton accumulation. Supplementing the rate of proton accumulation through CEF1 (A) or WWC (B) will increase pmf, the rate of proton efflux and, consequentially, the rate of ATP synthesis. However, since electrons on the reducing side of PSI return to the PQ/PQH2 pool via CEF1 or to water via WWC, NADPH output does not change. Since at steady-state, the rate of efflux will equal the rate of accumulation, pmf will also depend on how conductive the membranes are to proton flux. Thus, decreasing conductivity (C) will require an increase in pmf to balance proton accumulation with efflux. Since the steady-state rate of proton flux does not change in proportion to electron flux to NADPH, the relative outputs of ATP and NADPH remain constant. Finally, if, under most conditions the ΔpH partition is approximately 50% of pmf, collapsing the electric field component through counterion movements (D) would require an increase in ΔpH to sustain steady-state proton flux. In all cases, the sensitivity of qE to LEF (qE/LEF) increases.

Changes in the H+/e ratio for LEF

The Q-cycle is a catalytic mechanism which couples electron transfer through the cyt b6f complex to the translocation of protons from the stroma to the lumen (reviewed by Kurisu et al., 2003; Sacksteder et al., 2000). For each electron transferred through LEF, one proton is released into the lumen from water oxidation, and one proton is taken up during PQ reduction at the QB site of PSII and released when PQH2 is oxidized at the Qo site of the cyt b6f complex. An additional proton is translocated by the Q-cycle, making the overall H+/e stoichiometry for LEF 3. Although several authors have proposed that the Q-cycle is facultative (reviewed by Berry and Rumberg, 1999; Cornic et al., 2000; Ivanov, 1993; Kramer and Crofts, 1993; Sacksteder et al., 2000), disengaging it (see review by Sacksteder et al., 2000) would lower the H+/e ratio to 2, thereby lowering the pmf generated by LEF and, consequently, the ATP/NADPH output ratio and qE sensitivity.

On the other hand, in vitro mechanistic studies of the cyt b6f complex indicated that the Q-cycle was very likely obligatory (Kramer and Crofts, 1993; Rich, 1988). Furthermore, comparisons of estimated fluxes of protons with LEF and with electron flux through the cyt b6f complex in vivo suggested a constant H+/e ratio from low to saturating light intensities (Sacksteder et al., 2000). It was concluded that the Q-cycle is ptobably continuously engaged under normal, non-stressed photosynthetic conditions.

These arguments are bolstered by recent structural studies of the mitochondrial cyt bc1 complexes (Zhang et al., 1998) (which are hom*ologous to the chloroplast cyt b6f complex) and cyt b6f complexes (Kurisu et al., 2003; Stroebel et al., 2003) which have led to proposals that the ‘Rieske’ iron-sulphur protein ‘gates’ electron transfer by undergoing large-scale conformational changes during catalysis, essentially forcing the complex to shuttle protons via the Q-cycle (reviewed by Roberts et al., 2001; Zhang et al., 1998). It was concluded that there is a strong mechanistic basis for a constant H+/e ratio at the cyt b6f complex and that differential engagement of the Q-cycle probably does not account for ATP/NADPH balancing or for variable sensitivities of down-regulatory processes.

Alternate electron transfer cycles

Various light-driven cyclic electron transfer pathways have been proposed to translocate protons across the thylakoid and thus drive ATP production or initiate qE in the absence of NADP+ reduction. Two of the pathways, cyclic electron flow around PSI (CEF1) and the water-water cycle (WWC), have gained support in recent years and are discussed here.

Cyclic electron flux around PSI

CEF1 bypasses the photosynthetic Z-scheme by involving only one of the two photosystems, PSI. Light excites PSI, resulting in reduction of its FeS centres and oxidation of its primary chlorophyll donor, P700. Just as in LEF, the oxidized

\(\mathrm{P}_{700}^{{+}}\)

is reduced by electrons from the PQ pool, via the cyt b6f complex and PC (Bendall and Manasse, 1995). Electrons on the reducing side of PSI eventually reduce PQ, completing the cycle. There is no net reduction of Fd or NADP+ but flux of electrons through the cycle will translocate protons to the lumen, resulting in pmf, which can drive ATP synthesis and activate qE.

At least four pathways have been proposed to link the reducing side of PSI with the PQ pool (Fig. 1B, paths 1–4). First, a linkage may occur via a ferredoxin-PQ oxidoreductase (FQR) (path 1), a pathway that has been shown to be sensitive to antimycin A (Bendall and Manasse, 1995). Recently, Shikanai and coworkers (Munekage et al., 2002) identified an Arabidopsis mutant, pgr5, lacking antimycin A-sensitive Fd reduction of the PQ pool, preliminary evidence that the PGR5 gene product may be involved in FQR-mediated CEF1. Second, an enzyme hom*ologous to complex I of mitochondria and bacteria (Edwards and Walker, 1983; Kubicki et al., 1996), NAD(P)H-PQ oxidoreductase (NDH) (path 2), has been suggested to be involved in CEF1, but its role in vivo remains ambiguous, as suggested by deletion studies under normal (Endo et al., 1999; Horvath et al., 2000) and stress conditions (Barth and Krause, 2002; Sazanov et al., 1998). However, evidence presented by Shikanai and coworkers (Munekage et al., 2004) suggests that paths 1 and 2 act in parallel. Third, Cramer and co-workers (Zhang et al., 2001) suggested that, based on the copurification of FNR with the b6f complex, an FNR/b6f super complex (path 3) may operate in a third type of CEF1, a pathway that has been verified in vitro (Zhang et al., 2001) but the details of which remain unresolved (Kramer, 1990). Lastly, based on recent structures of the b6f complex (Kurisu et al., 2003; Stroebel et al., 2003), an unexpected haem group in between bH and the stromal phase has been identified, hinting at a role for this extra haem in mediating electron transfer from the reducing side of PSI (path 4), although a physiological pathway has yet to be identified (Stroebel et al., 2003). While several potential PQ reduction pathways have been identified, none can be established as dominant, nor can specific roles for the individual pathways be identified. Indeed, it is possible that the PQ reductase activities serve functions other than photosynthetic (Sazanov et al., 1998).

In vivo estimates of CEF1 rates

Care must be taken before accepting in vitro rates as reflecting those that can occur in vivo, especially since CEF1 is known to be well-regulated and a measurable change in its relative rate may appear only under special conditions. There is strong evidence for participation of CEF1 in ATP synthesis in green algae (e.g. Chlamydomonas) and cyanobacteria (Depege et al., 2003; Finazzi et al., 2002), as well as in C4 plant bundle sheath chloroplasts (Kubicki et al., 1996). However, the situation in C3 vascular plants is clearly unresolved, with the bulk of the evidence pointing to only minor contributions of CEF1 under steady-state conditions.

Several groups have estimated CEF1 rates in C3 vascular plants under steady-state conditions. These measurements are difficult because the electrons flow in a cycle, and no readily measurable, stable products are formed. One approach to indicate the activation of CEF1 is to estimate steady-state transthylakoid ΔpH using pH-indicator dyes, or the onset of qE with LEF. The argument is that at a given LEF, CEF1 will increase pmf, thereby decreasing lumen pH, and thus increasing qE (Cornic and Briantais, 1991; Heber, 2002). However, it is argued below that such effects can equally result from the engagement of Type II mechanisms, which have been shown to alter the relationship between LEF and steady-state pmf, as well as between pmf and qE.

Most commonly, the relative fluxes of electrons through different parts of the electron transfer chain are compared to estimate the relative engagements of LEF and CEF1. In steady-state LEF, the rates of electron transfer through PSII should equal that through PSI (Genty et al., 1990; Klughammer and Schreiber, 1994; Kramer and Crofts, 1996; Ort and Baker, 2002) or the cyt b6f complex (Klughammer and Schreiber, 1994; Sacksteder and Kramer, 2000). The engagement of CEF1 should increase electron flux through PSI over that through PSII. Likewise, the ratio of proton translocation to LEF should increase with the engagement of CEF1 (Sacksteder et al., 2000). The fraction of overall photosynthetic energy storage attributable to PSII will change with the engagement of CEF1 (Herbert et al., 1990). Unfortunately, each of these techniques measures CEF1 only as a fraction of LEF and is only sensitive to changes in the ratio of CEF1:LEF (Bendall and Manasse, 1995; Kramer and Crofts, 1996), and low rates are not readily detected. A number of studies using such assays have found little evidence for changes in the fractional turnover of CEF1 in steady states as conditions were altered, and thus the general consensus appears to be that, in C3 vascular plants, CEF1 is either negligible or a fairly constant fraction of steady-state LEF (Genty et al., 1990; Herbert et al., 1990; Klughammer and Schreiber, 1994; Kramer and Crofts, 1996; Ort and Baker, 2002; Sacksteder and Kramer, 2000). On the other hand, in more recent papers other groups have reported substantial rates of CEF1 (15–100% of LEF) during photosynthetic induction (Joliot and Joliot, 2002) or anaerobiosis (Joet et al., 2001) or under high light, low temperature conditions (Clarke and Johnson, 2001) or drought stress (Golding and Johnson, 2003).

The water-water cycle (WWC)

In the WWC, electrons extracted from H2O by PSII are transferred through the ETC to PSI, where O2 acts as the terminal acceptor forming superoxide (

\(\mathrm{O}_{2}^{{-}},\)

Fig. 1B, WWC).

\(\mathrm{O}_{2}^{{-}}\)

is dismutated to hydrogen peroxide and dioxygen, a reaction that is catalysed by superoxide dismutase (SOD), and the hydrogen peroxide is reduced to H2O by ascorbate peroxidase, thereby completing the cycle. Since O2 is reduced more slowly by Fd than FNR, the WWC has been proposed to operate to a higher extent when concentrations of NADP+ are low. Although the WWC produces no net reductant, it does generate pmf, which may serve to drive ATP synthesis or to initiate down-regulation (Asada, 1996).

Because it shares nearly all reactions with LEF, the WWC is very difficult to distinguish from LEF (Heber, 2002) and thus it is not surprising that issues concerning the relative contribution of the WWC to overall electron transfer have not yet been resolved. Much of the literature (Foyer and Noctor, 2000; Heber, 2002) suggests that, in vivo, the WWC is a relatively minor contributor to LEF. An estimate based on a survey of more recent work (Badger et al., 2000) suggests that, at most, WWC operates at 10% of LEF of C3 photosynthesis, even under conditions of extreme stress. Moreover, others have observed little to no WWC under conditions that should favour NADPH accumulation, such as lowered RUBISCO levels (Ruuska et al., 2000) or low temperatures (Clarke and Johnson, 2001).

By contrast, higher flux capacities for WWC have been observed in isolated chloroplasts of C3 plants (Backhausen et al., 2000; Badger et al., 2000), suggesting that conditions which favour WWC may not be simple to produce in vivo. However, there is evidence for the active engagement of the WWC in conjunction with CEF1 in rice leaves, during photosynthetic induction (Makino et al., 2002). It was suggested that the supplemental proton flux was required to generate additional ATP for the initiation of the Calvin–Benson cycle from a dark-adapted state. Furthermore, suppressed expression of thylakoid-associated Cu/Zn-SOD in Arabidopsis suppressed photosynthetic activity and growth, which is consistent with the need for detoxification of

\(\mathrm{O}_{2}^{{-}}\)

generated by photosynthesis (Rizhsky et al., 2003). While this observation supports the presence of the WWC in vivo, it does not necessarily support a role for the WWC supplementing pmf during steady-state photosynthesis.

Type II: Flexibility without altering ATP/NADPH output ratio

While Type I mechanisms could be modulators of qE, effective engagement would require them to comprise a large fraction of total electron flux, leading to mismatch in supply and demand for ATP and NADPH. By contrast, Type II mechanisms, as depicted in Fig. 1, allow the regulation of qE sensitivity without perturbing the ATP/NADPH ratio.

Alteration of qE response to lumen pH

One way to alter qE sensitivity would be to change the response capacity of qE to lumen pH. Over developmental time-scales, the differential accumulation of antenna and xanthophyll components has been shown to alter qE sensitivity (Demmig-Adams and Adams III, 1996). Hypothetically, more dynamic changes in qE sensitivity could occur though alterations in the pH response of the molecular components of qE. For example, covalent modification of VDE or PsbS could shift either pH dependence of VDE or pKas of protonatable groups on PsbS, respectively. Alternatively, components in the membrane could be modified, affecting the propensity of LHCs to aggregate or associate with the xanthophyll components, processes which have been linked to exciton dissipation by qE (reviewed by Horton et al., 1996). The predicted outcome, in all cases, would be a range of sensitivities of qE to ΔpH. However, in tobacco, a constant relationship was observed between qE and estimates of light-driven pmf changes, over conditions where qE sensitivity was substantially altered by changing CO2 levels (Kanazawa and Kramer, 2002), while under extreme acceptor limiting conditions qE was a continuous function of ΔpH (Avenson et al., 2004). These observations suggest that a constant relationship exists between lumen pH and qE and that modifications in antenna response do not account for short-term changes in qE sensitivity, under these conditions.

The importance of pmf composition for modulating qE response

Since qE is triggered by the ΔpH, but not the Δψ (electric field) component of thylakoid pmf, one way to change qE sensitivity would be to alter the manner in which pmf is stored. The chemisomotic mechanism, first described by Peter Mitchell, states that pmf is thermodynamically composed of the sum of the ΔpH and Δψ potentials (Mitchell, 1966). Many of the earlier characterizations of pmf were performed by monitoring ATP synthesis in intact thylakoids as a function of ΔpH produced by pH jump and/or by measuring ΔpH-dependent uptake of radiolabelled or fluorescent amines (Davenport and McCarty, 1986; Junesch and Gräber, 1985; Schuldiner et al., 1972). While useful for defining the thresholds of activation and other energetic parameters, these studies ignored and actively suppressed the Δψ component of pmf through the use of uncouplers and/or relatively high concentrations of counterions. Direct measurements of Δψ, made using salt-filled microelectrodes (Vredenberg and Tonk, 1975), helped to popularize the notion that it contributed little or negligibly to steady-state pmf, despite changes observed in vivo in the electrochromic shift (ECS) (Finazzi and Rappaport, 1998; Joliot and Joliot, 1989; Sacksteder et al., 2000) or measurement of Δψ-dependent ATP synthesis (Hangarter and Good, 1982; Junesch and Gräber, 1991), which suggested the contrary. Lately, it has been argued that under permissive conditions, it is unlikely that ΔpH solely comprises pmf (reviewed in Cruz et al., 2001; Kramer et al., 1999). In essence, a ΔpH requirement of 2–3 to activate ATP synthesis (Kramer and Crofts, 1989) yields a lumen pH that is inconsistent with the pH sensitivities of PSII and PC and with the pH-dependent rates of VDE and cyt b6f, observed in vivo.

In much of the authors' recent work, the ECS has been exploited as an endogenous probe for changes in transthylakoid Δψ during light-to-dark transitions (Avenson et al., 2004; Cruz et al., 2001; Kanazawa and Kramer, 2002; Sacksteder et al., 2000). The relevance of the ECS to pmf was first reported by Junge and Witt (1968). ECS refers to a Δψ-induced ‘shift’ in the absorption spectrum of pigments (i.e. chlorophyll and carotenoids) embedded in the thylakoid membrane. The peak of the difference spectrum occurs at 515–520 nm and has been shown to be a linear indicator of the strength of the transthylakoid Δψ (Witt and Zickler, 1973). One particular advantage of using the ECS is that it is non-invasive, allowing in vivo measurements on intact leaves. Generally, two techniques were employed when using ECS to probe pmf, both of which are variations of Dark Interval Relaxation Kinetic (DIRK) analysis (Sacksteder and Kramer, 2000). The DIRK technique uses brief (<500 ms) dark intervals to create reproducible perturbations in steady-state electron and proton fluxes. The initial rate of the ECS decay has been attributed to proton flux through the ATP synthase (Kramer and Crofts, 1989), and initial rates have been argued to reflect steady-state LEF or proton flux linearly (Sacksteder et al., 2000). From steady-state conditions, these rapid ECS decay kinetic traces are fit to mono-exponential decays (Fig. 3A), giving decay times (τECS) inversely proportional to proton conductivity (

\(g_{\mathrm{H}^{{+}}}\)

⁠) across the membrane (i.e. predominantly through the ATP synthase). The full extent of the decay (ECSt) should be proportional to the light-induced, steady-state pmf (Avenson et al., 2004; Kanazawa and Kramer, 2002; Sacksteder and Kramer, 2000).

Fig. 3.

Plasticity in light reactions of photosynthesis for energy production and photoprotection (5)

Open in new tabDownload slide

DIRK analysis of the ECS estimates pmf and its relative partitioning into Δψ and ΔpH. Tobacco leaves from intact plants were clamped into the measuring chamber of the spectrophotometer, which was purged with a stream of water-saturated air. Following 15 min of actinic illumination with 520 μmol of photons m−2 s−1 red light, ECS decay kinetic traces were collected for 300 ms (A) and 120 s (B) dark intervals. Over the 300 ms time-scale the ECS signal predominates and the extent of the decay estimates light-induced pmf, or ECSt. The time constant for decay of the signal (τECS), estimated from a first order fit (red line) of the data (r2=0.99) is inversely proportional to the proton conductivity of the ATP synthase (

\(g_{\mathrm{H}^{{+}}}\)

⁠). Over longer periods of analysis (i.e. minutes), light scattering contributes more significantly to apparent absorbance changes at 520 nm, necessitating deconvolution of the ECS kinetic trace (as described in Kramer and Sacksteder, 1998). From the deconvoluted trace, arbitrary amplitudes for Δψ (ECSss) and ΔpH (ECSinv) may be derived from the extents of the decay from steady-state to the dark baseline (ECSss) and from the dark baseline to full extent of the ECS inversion (ECSinv), respectively. (C) depicts the ‘ohmic’ relationship between force (pmf) and flux (

\(v_{\mathrm{H}^{{+}}}\)

). Flux increases linearly with force, with a slope equal to the conductivity,

\(g_{\mathrm{H}^{{+}}},\)

when the force is above the threshold of activation (arrows). With tight coupling of flux and ATP synthesis,

\(v_{\mathrm{H}^{{+}}}\)

is equivalent to the H+/ATP ratio, n, multiplied by the ATP synthesis rate. Increasing the threshold of activation (as happens with oxidation of the ATP synthase γ disulphide (Junesch and Gräber, 1985) or with an increase in ΔGATP (Hangarter and Good, 1982)) does not change the conductivity above activation (dotted line).

With longer dark intervals (i.e. minutes), following the rapid, initial decay the ECS relaxes to a dark level (Fig. 3B). Determination of the dark baseline allows tentative separation of Δψ- and ΔpH-driven decays of pmf (Cruz et al., 2001). Since the ECS is a linear indicator of Δψ, the extent of its decay from steady-state to baseline (ECSss) should be proportional to the light-induced Δψ. However, the ECS continues to decay below baseline, indicating ‘inversion’ of the electric field with respect to steady-state levels. This effect arises from the continued efflux of protons from the lumen, driven by ΔpH (and beyond that driven by Δψ alone). Since the ECS decay should continue until ΔpH essentially reaches equilibrium with the inverted Δψ, the extent to which ECS drops below baseline (ECSinv) should be proportional to the light-induced ΔpH, at least under appropriate conditions (Cruz et al., 2001). The sum of the amplitudes for ECSss and ECSinv (i.e. ECSt) should be proportional to the light-induced pmf. Using this ‘partition analysis’, the fraction of pmf stored as Δψ (or ΔpH) may be expressed as the ECSss (or ECSinv) divided by ECSt.

Modulation of pmf by proton conductivity

Above the threshold of activation for ATP synthesis, a linear relationship exists between pmf and proton efflux until the ATP synthase pool reaches the maximum rate of turnover (Hangarter and Good, 1982; Junesch and Gräber, 1985; Kramer and Crofts, 1989). The slope of this relationship,

\(g_{\mathrm{H}^{{+}}},\)

is a measure of the response of transthylakoid efflux to changes in driving force (Fig. 3C). It is important to note that changing

\(g_{\mathrm{H}^{{+}}}\)

will not change the steady-state rate of H+ efflux at a given LEF, but it will change the pmf required to sustain this rate (Fig. 2). Since qE should be a continuous function of pmf, provided that ΔpH is a constant fraction of pmf, qE sensitivity could be modulated by changes in

\(g_{\mathrm{H}^{{+}}}.\)

Indeed, correlative changes in qE sensitivity and

\(g_{\mathrm{H}^{{+}}}\)

have been observed in tobacco. Through DIRK analysis of the ECS, large decreases in

\(g_{\mathrm{H}^{{+}}}\)

were observed as CO2 levels were lowered from 2000 to nearly 0 ppm with coincident increases in qE sensitivity (Kanazawa et al., 2001; Kanazawa and Kramer, 2002). Further analysis suggested that essentially all changes in qE sensitivity could be explained by modulation of

\(g_{\mathrm{H}^{{+}}}\)

alone.

Variable partitioning of pmf

Variable pmf partitioning allows a more flexible relationship to exist between steady-state ATP synthesis and qE. For example, increasing the fraction ΔpH at a given pmf will increase the qE response without affecting the ATP synthetic rate (Fig. 2). Such effects would be important in cases where the physiological status of the leaf requires qE to be large, but LEF is too low to support a large enough pmf. Variable pmf partitioning will lead to discontinuities in relationships between pmf and qE as well as LEF and qE, and recently, this effect was observed in tobacco leaves under low O2 and low CO2 (Avenson et al., 2004). Partition analysis of the ECS kinetics indicated a relative increase in the fraction of pmf stored as ΔpH, suggesting that variable partitioning of pmf was responsible for enhancing the sensitivity of qE. Furthermore, decreases in

\(g_{\mathrm{H}^{{+}}}\)

were observed as well, suggesting that variable pmf partitioning and

\(g_{\mathrm{H}^{{+}}}\)

modulation may act in concert to increase qE sensitivity.

Previous work (Cruz et al., 2001) suggested that, if the thylakoid membrane contains only passive channels to allow counterions to dissipate Δψ, the extent to which pmf will be stored as Δψ and ΔpH will depend largely on the proton buffering capacity of the lumen and the concentration of counterions. The collective work of several groups suggest that buffering capacity will depend on the concentration of fixed buffering groups (Ewy and Dilley, 2000; Junge et al., 1979; van Kooten et al., 1986) and that in vivo this is unlikely to change to a large extent (Junge et al., 1979; van Kooten et al., 1986). Thus, postulating the existence of active ion transport mechanisms in the thylakoid (e.g. coporters, antiporters, ion-transporting pumps), the most likely mechanism to alter pmf partitioning is the concentration of counterions, and it is not difficult to imagine that these could be regulated in some way in vivo to adjust qE sensitivity.

An integrated model for photosynthetic flexibility

As discussed earlier, without Type I flexibility, LEF would yield a ratio of 1.3 ATP/NADPH (Allen, 2002). As estimated from the combined energy requirements for CO2 fixation under photorespiring conditions and nitrate assimilation to glutamate, a ratio of ∼1.43 ATP/NADPH (Edwards and Walker, 1983) would be needed through photochemistry in the chloroplast, yielding a deficit of about 0.13 ATP per NADPH. The net contributions of Type I mechanisms to increasing relative ATP output will depend on their H+/e coupling stoichiometries. The WWC, which uses essentially the same reactions as LEF, probably produces an H+/e of 3. In the case of CEF1, the H+/e will be partly determined by the pathway for PQ reduction, which is not well understood, especially in C3 vascular plants. Conservative estimates, based on various models, place the lower-upper bounds for H+/e between 2 and 4. Using these values, the WWC would need to run at a rate of about 12%, or CEF1 between 18% and 9%, that of LEF to fill the ATP deficit between the light reactions and downstream metabolism. While current estimates of WWC and CEF1 capacity are close to and in some cases exceed this requirement, as noted by Makino et al. (2002), to balance output it is probable that they run in concert, possibly with other electron sinks such as the malate valve (Fridlyand et al., 1998) or chlororespiration (Cardol et al., 2003) or with mitochondrial respiration (Noctor and Foyer, 1998). Although this contribution may seem small, the impact of Type I mechanisms on C3 photosynthesis might be quite significant. Indeed, Arabidopsis double mutants for PGR5 and NDH show substantial decreases in capacity for NPQ, probably due to the observed decreases in LEF (and subsequent pmf formation), created by an uncorrected imbalance in the supply and demand ratios for ATP and NADPH, leading to metabolic conjestion (Munekage et al., 2004). Thus, these data are consistent with the view presented here, in that the relatively small contributions indirectly affect proton translocation.

Assignment of this specific role to Type I mechanisms is supported by the observed evolutionary adaptations of the photosynthetic apparatus. For example, if carbon fixation is looked at specifically, there is considerable evidence that steady-state CEF1 is small in C3 vascular plants, where the expected output balance of LEF is close to, but does not precisely match, biochemical demand. The exception is induction of photosynthesis when priming of the carbon cycle might also require additional ATP (Poolman et al., 2003), as inferred from high CEF1 rates reported by Joliot and Joliot (2002) and Cardol et al. (Cardol et al., 2003). However, some C4 plants require a ratio of ATP/NADPH of 5:2. In species like maize and sorghum, mesophyll chloroplasts generate most, if not all, of the reductive power while bundle sheath chloroplasts function to produce ATP, likely via a CEF1 pathway (Edwards and Walker, 1983; Ivanov et al., 2001). Similarly, cyanobacteria and green algae, for which robust CEF1 is well-documented, need a higher PSI/PSII ratio to fix CO2 than do C3 vascular plants, in part because they possess ATP-driven CO2 concentrating mechanisms (Ogawa and Kaplan, 2003; Turpin and Bruce, 1990).

However, extremes in acceptor limitation (namely O2 and CO2) can induce rather dramatic (up to ∼6-fold) changes in the sensitivity of qE to LEF (Avenson et al., 2004; Kanazawa and Kramer, 2002). To account for such a robust response by themselves, CEF1 or WWC would have to occur at rates about 5-fold larger than that of LEF. Even the largest estimates of CEF1 and WWC capacity fall far short of this. Thus, it is argued that CEF1 cannot by itself account for the observed large changes in qE sensitivity to LEF, and that contributions from Type II mechanisms are therefore necessary.

It is important to keep in mind that all protons translocated into the lumen, either via LEF or a cyclic process, pass back across the thylakoid membrane, mainly through the ATP synthase. This implies that any increase in pmf generation by CEF1 or WWC (Type I mechanisms) will result in additional ATP synthesis, as long as uncoupling or ‘slip’ in the ATP synthase reaction is negligible, as shown by Junge and coworkers (Groth and Junge, 1993). Moreover, the augmented pmf should also increase regulatory sensitivity via qE, as long as static pmf partitioning etc., accompany changes in Type I mechanisms. However, as discussed previously, large increases in Type I flux could result in an excessive increase in the supply ratio of ATP/NADPH. While, in principle, the problem could be solved by dissipating ATP non-productively in a futile cycle, no such futile ATPase activity has yet been identified. By contrast, Type II mechanisms increase regulatory sensitivity without altering ATP/NAPDH output ratios, and rather strong evidence has been presented that when an increase in qE sensitivity is all that is needed, for example, under LEF-limited conditions, Type II mechanisms are activated.

These observations form the basis of an integrated model, where Type I mechanisms (CEF1 and WWC) provide plasticity at the level of ATP/NADPH, no doubt also impacting qE sensitivity, while Type II mechanisms play a more significant role in adjusting qE sensitivity without altering the ATP/NADPH output ratio. Furthermore, these mechanisms are not mutually exclusive and may act separately or in parallel to modulate qE sensitivity.

One implication of the integrated model is that, for both types of mechanisms, regulation will be mediated through metabolic pools of the reactants or products of the light reactions. For example, for Type I mechanisms to operate effectively, the levels of their induction would need to be dictated by fluctuations in ATP consumption relative to NADPH. One possible model is that induction will be sensitive to the redox poise of the NADP+/NADPH couple or the intermediate carriers of the ETC. Indeed, there is strong evidence that CEF1 must be properly redox poised to operate in vascular plants (Bendall and Manasse, 1995; Joet et al., 2001; Joliot and Joliot, 2002). Moreover, in Chlamydomonas, state transitions, which have a large effect on excitation energy distribution between PSI and PSII (Delosme et al., 1996), appear to trigger CEF1 (Depege et al., 2003; Finazzi et al., 2002), and this process is regulated by the phosphorylation of antenna complexes and is initiated by changes in the redox state of the PQ pool (reviewed in Allen and Forsberg, 2001; Haldrup et al., 2001).

Similarly, engagement of Type II mechanisms would be expected to be linked to proportionate changes in overall flux of NADPH and ATP. In fact, these mechanisms appear to be induced under conditions where LEF and ATP synthesis are limited by the availability of electron acceptors and Pi acceptors (e.g. low CO2). This fits adequately with a proposed model where

\(g_{\mathrm{H}^{{+}}}\)

is modulated by stromal Pi levels (Kanazawa and Kramer, 2002). Pi levels in chloroplasts are typically 20 mM and may drop to 10 mM during photosynthesis (Furbank et al., 1987; Usuda, 1988). Under stress conditions Pi concentrations may dip as low as 2 mM (Sharkey and Vanderveer, 1989). Since 1–2 mM of stromal Pi is inactive (Furbank et al., 1987; Robinson and Giersch, 1987), and not available for photophosphorylation, active Pi concentration could be close to the reported Km, 0.6 mM (Selman and Selman-Reimer, 1981). Indeed, decreases in the

\(g_{\mathrm{H}^{{+}}}\)

of spinach and Arabidopsis leaf discs have been observed when pretreated with mannose, which has been shown to act in vivo as a phosphate sink. Commensurate increases in pmf and qE were also observed, and all effects were reversed with phosphate replenishment (K Takizawa, unpublished data). An alternative possibility is that ATP synthase activity is modulated allosterically. Interestingly, evidence has been presented that a 14-3-3 protein can bind a phosphorylated chloroplast ATP synthase and that binding partially inhibits turnover (Bunney et al., 2001).

Abbreviations: CEF1, cyclic electron flow around PSI; cyt, cytochrome; ΔpH, transthylakoid pH gradient; Δψ, transthylakoid electric field; DIRK, Dark Interval Relaxation Kinetics; ECS, electrochromic shift; ECSinv, inverted ECS; ECSss, steady-state ECS; ECSt, full extent of the ECS decay; ETC, electron transfer chain; Fd, ferredoxin; FNR, ferredoxin:NADP+ oxidoreductase;

\(g_{\mathrm{H}^{{+}}},\)

proton conductivity; H+/e, proton to electron stoichiometry; LEF, linear electron flux; LHCs, light-harvesting complexes; NPQ, non-photochemical exciton quenching; OEC, oxygen evolving complex; PC, plastocyanin; pmf, proton motive force; PQ, plastoquinone; PS, photosystem; qE, energy-dependent quenching; SOD, superoxide dismutase; τECS, ECS decay time; VDE, violaxanthin deepoxidase; WWC, water–water cycle.

References

Allen JF.

2002

. Photosynthesis of ATP—electrons, proton pumps, rotors, and poise.

Cell

110

,

273

–276.

Allen JF, Forsberg J.

2001

. Molecular recognition in thylakoid structure and function.

Trends in Plant Science

6

,

317

–326.

Anderson B, Barber J.

1996

. Mechanisms of photodamage and protein degradation during photoinhibition of photosystem II. In: Baker NR, ed. Photosynthesis and the environment. Dordrecht, The Netherlands: Kluwer Academic Publishers, 101–121.

Anderson JM, Park Y-I, Chow WS.

1997

. Photoinactivation and photoprotection of photosystem II in nature.

Physiologia Plantarum

100

,

214

–223.

Asada K.

1996

. Radical production and scavenging in the chloroplasts. In: Baker NR, ed. Photosynthesis and the environment. Dordrecht, The Netherlands: Kluwer Academic Publishers, 123–150.

Avenson T, Cruz JA, Kramer D.

2004

. Modulation of energy-dependent quenching of excitons (qE) in antenna of higher plants.

Proceedings of the National Academy of Sciences, USA

101

,

5530

–5535.

Backhausen JE, Kitzman C, Horton P, Scheibe R.

2000

. Electron acceptors in isolated intact spinach chloroplasts act heirarchically to prevent over-reduction and competition for electrons.

Photosynthesis Research

64

,

1

–13.

Badger MR, von Caemmerer S, Ruuska S, Nakano H.

2000

. Electron flow to oxygen in higher plants and algae: rates and control of direct photoreduction (Mehler reaction) and rubisco oxygenase.

Philosophical Transactions of the Royal Society of London, Series B: Biological Sciences

355

,

1433

–1446.

Barth C, Krause GH.

2002

. Study of tobacco transformants to assess the role of chloroplastic NAD(P)H dehydrogenase in photoprotection of photosystems I and II.

Planta

216

,

273

–279.

Bendall DS, Manasse RS.

1995

. Cyclic photophosphorylation and electron transport.

Biochimica et Biophysica Acta

1229

,

23

–38.

Berry S, Rumberg B.

1999

. Proton to electron stoichiometry in electron transport of spinach thylakoids.

Biochimica et Biophysica Acta

1410

,

248

–261.

Bunney TD, van Walraven HS, de Boer AH.

2001

. 14-3-3 protein is a regulator of the mitochondrial and chloroplast ATP synthase.

Proceedings of the National Academy of Sciences, USA

98

,

4249

–4254.

Cardol P, Gloire G, Havaux M, Remacle C, Matagne R, Franck F.

2003

. Photosynthesis and state transitions in mitochondrial mutants of Chlamydomonas reinhardtii affected in respiration.

Plant Physiology

133

,

2010

–2020.

Clarke JE, Johnson GN.

2001

. In vivo temperature dependence of cyclic and psuedocyclic electron transport in barley.

Planta

212

,

808

–816.

Cornic G, Briantais JM.

1991

. Partitioning of photosynthetic electron flow between CO2 and O2 reduction in a C3 leaf (Phaseolus vulgaris L.) at different CO2 concentrations and during drought stress.

Planta

183

,

178

–184.

Cornic G, Bukhov NG, Wiese C, Bligny R, Heber U.

2000

. Flexible coupling between light-dependent electron and vectorial proton transport in illuminated leaves of C3 plants. Role of photosystem I-dependent proton pumping.

Planta

210

,

468

–477.

Cruz JA, Sacksteder CA, Kanazawa A, Kramer DM.

2001

. Contribution of electric field (Δψ) to steady-state transthylakoid proton motive force in vitro and in vivo. Control of pmf parsing into Δψ and ΔpH by counterion fluxes.

Biochemistry

40

,

1226

–1237.

Davenport JW, McCarty RE.

1986

. Relationships between rates of steady-state ATP synthesis and the magnitude of the proton-activity gradient across thylakoid membranes.

Biochimica et Biophysica Acta

851

,

136

–145.

Delosme R, Olive J, Wollman F-A.

1996

. Changes in light energy distribution upon state transitions: an in vivo photoacoustic study of the wild type and photosynthesis mutants from Chlamydomonas reinhardtii.

Biochimica et Biophysica Acta

1273

,

150

–158.

Demmig-Adams B, Adams III WW.

1996

. The role of xanthophyll cycle carotenoids in the protection of photosynthesis.

Trends in Plant Science

1

,

21

–26.

Depege N, Bellafiore S, Rochaix JD.

2003

. Role of chloroplast protein kinase Stt7 in LHCII phosphorylation and state transition in Chlamydomonas.

Science

299

,

1572

–1575.

Diner BA, Babco*ck GT.

1996

. Structure, dynamics, and energy conversion efficiency in photosystem II. In: Ort DR, Yocum CF, eds. Oxygenic photosynthesis: the light reactions. The Netherlands: Kluwer Academic Publishers, 213–247.

Edwards GE, Walker DA.

1983

. C3, C4: mechanisms, and cellular and environmental regulation of photosynthesis. Textbook on C3, C4 photosynthesis. Oxford: Blackwell Scientific.

Endo T, Shikanai T, Takabayashi T, Asada K, Sato F.

1999

. The role of chloroplastic NAD(P)H dehydrogenase in photoprotection.

FEBS Letters

457

,

5

–8.

Eskling M, Emanuelsson A, Akerlund H-E.

2001

. Enzymes and mechanisms for violaxanthin-zeaxanthin conversion. In: Aro E-M, Anderson B, eds. Regulation of photosynthesis, Vol. 100. Dordrecht, The Netherlands: Kluwer Academic Publishers, 806–816.

Ewy RG, Dilley RA.

2000

. Distinguishing between luminal and localized proton buffering pools in thylakoid membranes.

Plant Physiology

122

,

583

–596.

Finazzi G, Rappaport F.

1998

. In vivo characterization of the electrochemical proton gradient generated in darkness in green algae and its kinetic effects on cytochrome b6f turnover.

Biochemistry

37

,

9999

–10005.

Finazzi G, Rappaport F, Furia A, Fleischmann M, Rochaix JD, Zito F, Forti G.

2002

. Involvement of state transitions in the switch between linear and cyclic electron flow in Chlamydomonas reinhardtii.

EMBO Reports

3

,

280

–285.

Foyer CH, Noctor G.

2000

. Oxygen processing in photosynthesis: regulation and signalling.

New Phytologist

146

,

359

–388.

Fridlyand LE, Backhausen JE, Scheiber R.

1998

. Flux control of the malate valve in leaf cells.

Archives of Biochemistry and Biophysics

349

,

290

–298.

Furbank RT, Foyer CH, Walker DA.

1987

. Regulation of photosynthesis in isolated spinach chloroplasts during orthophosphate limitation.

Biochimica et Biophysica Acta

894

,

552

–561.

Genty B, Harbinson J, Baker NR.

1990

. Relative quantum efficiencies of the two photosystems of leaves in photorespiratory and non-photorespiratory conditions.

Plant Physiology and Biochemistry

28

,

1

–10.

Golding AJ, Johnson GN.

2003

. Down-regulation of linear and activation of cyclic electron transport during drought.

Planta

218

,

107

–114.

Groth G, Junge W.

1993

. Proton slip of the chloroplast ATPase: its nucleotide dependence, energetic threshold, and relation to an alternating site mechanism of catalysis.

Biochemistry

32

,

8103

–8111.

Haldrup A, Jensen PE, Lunde C, Scheller HV.

2001

. Balance of power: a view of the mechanism of photosynthetic state transitions.

Trends in Plant Science

6

,

301

–305.

Hangarter RP, Good ND.

1982

. Energy thresholds for ATP synthesis in chloroplasts.

Biochimica et Biophysica Acta

681

,

396

–404.

Heber U.

2002

. Irrungen, wirrungen? The Mehler reaction in relation to cyclic electron transport in C3 plants.

Photosynthesis Research

73

,

223

–231.

Heber U, Walker D.

1992

. Concerning a dual function of coupled cyclic electron transport in leaves.

Plant Physiology

100

,

1621

–1626.

Herbert SK, Fork DC, Malkin S.

1990

. Photoacoustic measurements in vivo of energy storage by cyclic electron flow in algae and higher plants.

Plant Physiology

94

,

926

–934.

Horton P.

1989

. Interactions between electron transport and carbon assimilation: regulation of light harvesting and photochemistry. In: Briggs W, ed. Photosynthesis, plant biology, Vol. 9. New York: A. Liss Inc, 393–406.

Horton P, Ruban A, Walters R.

1996

. Regulation of light harvesting in green plants.

Annual Review of Plant Physiology and Plant Molecular Biology

47

,

655

–684.

Horton P, Ruban AV, Young AJ.

1999

. Regulation of the structure and function of the light-harvesting complexes of photosystem II by the xanthophyll cycle. In: Frank HA, Young AJ, Britton G, Cogdell RJ, eds. The photochemistry of carotenoids. Dordrecht, The Netherlands: Kluwer Academic Publishers, 271–291.

Horvath EM, Peter SO, Joet T, Rumeau D, Cournac L, Horvath GV, Kavanagh TA, Schafer C, Peltier G, Medgyesy P.

2000

. Targeted inactivation of the plastid ndhB gene in tobacco results in an enhanced sensitivity of photosynthesis to moderate stomatal closure.

Plant Physiology

123

,

1337

–1350.

Ivanov BN.

1993

. Stoichiometry of proton uptake by thylakoids during electron transport in chloroplasts. In: Abrol YP, Mohanty P, Govindjee, eds. Photosynthesis photoreactions to plant productivity. New Delhi: Oxford & IBH Publishing Co. Pvt. Ltd., 110–128.

Ivanov BN, Sacksteder CA, Kramer DM, Edwards GE.

2001

. Light induced ascorbate dependent electron transport and membrane energization in chloroplasts of bundle sheath cells of the C4 plant maize.

Archives of Biochemistry and Biophysics

385

,

145

–153.

Joet T, Cournac L, Horvath EM, Medgyesy P, Peltier G.

2001

. Increased sensitivity of photosynthesis to antimycin A induced by inactivation of the chloroplast ndhB gene. Evidence for a participation of the NADH-dehydrogenase complex to cyclic electron flow around photosystem I.

Plant Physiology

125

,

1919

–1929.

Joliot P, Joliot A.

1989

. Characterization of linear and quadratic electrochromic probes in Chlorella sorokiniana and Chlamydomonas reinhardtii.

Biochimica et Biophysica Acta

975

,

355

–360.

Joliot P, Joliot A.

2002

. Cyclic electron transfer in plant leaf.

Proceedings of the National Academy of Sciences, USA

99

,

10209

–10214.

Junesch U, Gräber P.

1985

. The rate of ATP synthesis as a function of ΔpH in normal and dithiothreitol-modified chloroplasts.

Biochimica et Biophysica Acta

809

,

429

–434.

Junesch U, Gräber P.

1991

. The rate of ATP-synthesis as a function of ΔpH and Δψ catalyzed by the active, reduced H+-ATPase from chloroplasts.

FEBS Letters

294

,

275

–278.

Junge W, Ausländer W, McGeer AJ, Runge T.

1979

. The buffering capacity of the internal phase of thykaloids and the magnitude of the pH changes inside under flashing light.

Biochimica et Biophysica Acta

546

,

121

–141.

Junge W, Witt HT.

1968

. On the ion transport system in photosynthesis: investigations on a molecular level.

Zeitschrift für Naturforschung

23b

,

244

–254.

Kanazawa A, Kiirats O, Edwards G, Cruz J, Kramer DM.

2001

. New steps in the regulation of photosynthesis: the influence of CF0-CF1 ATP synthase conductivity on the sensitivity of antenna down-regulation. Proceedings of the XIIth International Congress on Photosynthesis, Vol. S3-047. Collingwood, Vic. Australia: CSIRO Publishing, 1–4.

Kanazawa A, Kramer DM.

2002

. In vivo modulation of nonphotochemical exciton quenching (NPQ) by regulation of the chloroplast ATP synthase.

Proceedings of the National Academy of Sciences, USA

99

,

12789

–12794.

Klughammer C, Schreiber U.

1994

. An improved method, using saturating light pulses, for the determination of photosystem I quantum yield via

\(\mathrm{P}_{700}^{{+}}\)

absorbance changes at 830 nm.

Planta

192

,

261

–268.

Knaff DB.

1996

. Ferredoxin and ferredoxin-dependent enzymes. In: Ort DR, Yocum CF, eds. Oxygenic photosynthesis: the light reactions. The Netherlands: Kluwer Academic Publishers, 333–361.

Kramer DM.

1990

. Parallel instrumentational approaches to the study of the bioenergetics of photosynthesis in chloroplasts and intact plants. PhD thesis, Physiology and Biophysics, Urbana, Illinois: University of Illinois.

Kramer DM, Crofts AR.

1989

. Activation of the chloroplast ATPase measured by the electrochromic change in leaves of intact plants.

Biochimica et Biophysica Acta

976

,

28

–41.

Kramer DM, Crofts AR.

1993

. The concerted reduction of the high- and low-potential chains of the bf complex by plastoquinol.

Biochimica et Biophysica Acta

1183

,

72

–84.

Kramer DM, Crofts AR.

1996

. Control of photosynthesis and measurement of photosynthetic reactions in intact plants. In: Baker N, ed. Photosynthesis and the environment advances in photosynthesis. Dordrecht, The Netherlands: Kluwer Academic Press, 25–66.

Kramer DM, Cruz JA, Kanazawa A.

2003

. Balancing the central roles of the thylakoid proton gradient.

Trends in Plant Science

8

,

27

–32.

Kramer DM, Sacksteder CA.

1998

. A diffused-optics flash kinetic spectrophotometer (DOFS) for measurements of absorbance changes in intact plants in the steady-state.

Photosynthesis Research

56

,

103

–112.

Kramer DM, Sacksteder CA, Cruz JA.

1999

. How acidic is the lumen?

Photosynthesis Research

60

,

151

–163.

Kubicki A, Funk E, Westhoff P, Steinmueller K.

1996

. Differential expression of plastome-encoded ndh genes in the mesophyll and bundle sheath chloroplasts of the C4 plant Sorghum bicolor indicates that the complex I-hom*ologous NAD(P)H-plastoquinone oxidoreductase is involved in cyclic electron transport.

Planta

199

,

276

–281.

Kurisu G, Zhang H, Smith JL, Cramer WA.

2003

. Structure of the cytochrome b6f complex of oxygenic photosynthesis: tuning the cavity.

Science

302

,

1009

–1014.

Li XP, Gilmore AM, Caffarri S, Bassi R, Golan T, Kramer D, Niyogi KK.

2004

. Regulation of photosynthetic light harvesting involves intrathylakoid lumen pH sensing by the PsbS protein.

Journal of Biological Chemistry

279

,

22866

–22874.

Makino A, Miyake C, Yokota A.

2002

. Physiological functions of the water-water cycle (Mehler reaction) and the cyclic electron flow around PSI in rice leaves.

Plant and Cell Physiology

43

,

1017

–1026.

Malkin R.

1996

. Photosystem I electron transfer reactions-components and kinetics. In: Ort DR, Yocum CF, eds. Oxygenic photosyntheseis: the light reactions. The Netherlands: Kluwer Academic Press, 313–332.

Maxwell K, Johnson GN.

2000

. Chlorophyll fluorescence-a practical guide.

Journal of Experimental Botany

51

,

659

–668.

McCarty RE.

1996

. An overview of the function, composition and structure of the chloroplast ATP synthase. In: Ort DR, Yocum CF, eds. Oxygenic photosynthesis: the light reactions. The Netherlands: Kluwer Academic Publishers, 439–451.

Mitchell P.

1966

. Chemiosmotic coupling in oxidative and photosynthetic phosphorylation.

Biological Reviews of the Cambridge Philosophical Society

41

,

445

–502.

Müller P, Li X-P, Niyogi KK.

2001

. Non-photochemical quenching. A response to excess light energy.

Plant Physiology

125

,

1558

–1566.

Munekage V, Hashimoto M, Miyake C, Tomizawa K-I, Endo T, Tasaka M, Shikanai T.

2004

. Cyclic electron flow around photosystem I is essential for photosynthesis.

Nature

429

,

579

–582.

Munekage Y, Hojo M, Meurer J, Endo T, Tasaka M, Shikanai T.

2002

. PGR5 is involved in cyclic electron flow around photosystem I and is essential for photoprotection in Arabidopsis.

Cell

110

,

361

–371.

Noctor G, Foyer C.

1998

. Review article. A re-evaluation of the ATP:NADPH budget during C3 photosynthesis: a contribution from nitrate assimilation and its associated respiratory activity?

Journal of Experimental Botany

49

,

1895

–1908.

Ogawa T, Kaplan A.

2003

. Inorganic carbon acquisition systems in cyanobacteria.

Photosynthesis Research

77

,

105

–115.

Ort DR, Baker NR.

2002

. A photoprotective role for O2 as an alternative electron sink in photosynthesis?

Current Opinion in Plant Biology

5

,

193

–198.

Ort DR, Yocum CF.

1996

. Light reactions of oxygenic photosynthesis. In: Ort DR, Yocum CF, eds. Oxygenic photosynthesis: the light reactions. The Netherlands: Kluwer Academic Publishers, 1–9.

Owens TG.

1996

. Processing of excitation energy by antenna pigments. In: Baker N, ed. Photosynthesis and the environment. Dordrecht, The Netherlands: Kluwer Academic Publishers, 1–23.

Poolman M, Fell D, Raines C.

2003

. Elementary modes analysis of photosynthate metabolism in chloroplast stroma.

European Journal of Biochemistry

270

,

430

–439.

Rich PR.

1988

. A critical examination of the supposed variable proton stoichiometry of the chloroplast cytochrome bf complex.

Biochimica et Biophysica Acta

932

,

33

–42.

Rizhsky L, Liang H, Mittler R.

2003

. The water-water cycle is essential for chloroplast protection in the absence of stress.

Journal of Biological Chemistry

278

,

38921

–38925.

Roberts AG, Bowman MK, Kramer DM.

2001

. Certain metal ions are inhibitors of cytochrome b6f ‘Rieske’ iron-sulfur protein domain movements.

Biochemistry

41

,

4070

–4079.

Robinson SP, Giersch C.

1987

. Inorganic phosphate concentration in the stroma of isolated chloroplasts and its influence on photosynthesis.

Australian Journal of Plant Physiology

14

,

451

–462.

Ruuska S, Badger M, Andrews T, von Caemmerer S.

2000

. Photosynthetic electron sinks in transgenic tobacco with reduced amounts of Rubisco: little evidence for significant Mehler reaction.

Journal of Experimental Botany

51

,

357

–368.

Sacksteder CA, Kanazawa A, Jacoby ME, Kramer DM.

2000

. The proton to electron stoichiometry of steady-state photosynthesis in living plants: a proton-pumping Q-cycle is continuously engaged.

Proceedings of the National Academy of Sciences, USA

97

,

14283

–14288.

Sacksteder CA, Kramer DM.

2000

. Dark interval relaxation kinetics of absorbance changes as a quantitative probe of steady-state electron transfer.

Photosynthesis Research

66

,

145

–158.

Sazanov LA, Burrows PA, Nixon PJ.

1998

. The chloroplast Ndh complex mediates the dark reduction of the plastoquinone pool in response to heat stress in tobacco leaves.

FEBS Letters

429

,

115

–118.

Schuldiner S, Rottenberg H, Avron M.

1972

. Determination of ΔpH in chloroplasts.

European Journal of Biochemistry

25

,

64

–70.

Selman BR, Selman-Reimer S.

1981

. The steady-state kinetics of photophosphorylation.

Journal of Biological Chemistry

256

,

1722

–1726.

Sharkey TD, Vanderveer PJ.

1989

. Stromal phosphate concentration is low during feedback limited photosynthesis.

Plant Physiology

91

,

679

–684.

Stroebel D, Choquet Y, Popot JL, Picot D.

2003

. An atypical haem in the cytochrome b(6)f complex.

Nature

426

,

413

–418.

Turpin DH, Bruce D.

1990

. Regulation of photosynthetic light harvesting by nitrogen assimilation in the green alga Selenastrum minutum.

FEBS Letters

263

,

99

–103.

Usuda H.

1988

. Adenine nucleotide levels, the redox state of the NADP system, and assimilatory force in nonaqueously purified mesophyll chloroplasts from maize leaves under different light intensities.

Plant Physiology

88

,

1461

–1468.

van Kooten O, Snel JFH, Vredenberg WJ.

1986

. Photosynthetic free energy transduction related to the electric potential changes across the thylakoid membrane.

Photosynthesis Research

9

,

211

–227.

Vredenberg WJ, Tonk WJM.

1975

. On the steady-state electrical potential difference across the thylakoid membranes of chloroplasts in illuminated plant cells.

Biochimica et Biophysica Acta

387

,

580

–587.

Witt HT, Zickler A.

1973

. Electrical evidence for the field indicating absorption change in bioenergetic membranes.

FEBS Letters

37

,

307

–10.

Yamamoto HY, Bassi R.

1996

. Carotenoids: localization and function. In: Ort DR, Yocum CF, eds. Oxygenic photosynthesis: the light reactions. The Netherlands: Kluwer Academic Publishers, 539–563.

Zhang H, Whitelegge JP, Cramer WA.

2001

. Ferredoxin:NADP+ oxidoreductase is a subunit of the chloroplast cytochrome b6f complex.

Journal of Biological Chemistry

276

,

38159

–38165.

Zhang Z, Huang L, Shulmeister V, Chi Y, Kim K, Hung L, Crofts A, Berry E, Kim S.

1998

. Electron transfer by domain movement in cytochrome bc1.

Nature

392

,

677

–684.

Journal of Experimental Botany, Vol. 56, No. 411, © Society for Experimental Biology 2004; all rights reserved

Issue Section:

Regulation of Photosynthesis under Stress

Download all slides

Comments

0 Comments

Comments (0)

Submit a comment

You have entered an invalid code

Thank you for submitting a comment on this article. Your comment will be reviewed and published at the journal's discretion. Please check for further notifications by email.

Advertisem*nt

Citations

Views

11,304

Altmetric

More metrics information

Metrics

Total Views 11,304

9,409 Pageviews

1,895 PDF Downloads

Since 1/1/2017

Month: Total Views:
January 2017 17
February 2017 28
March 2017 26
April 2017 25
May 2017 33
June 2017 25
July 2017 14
August 2017 13
September 2017 16
October 2017 31
November 2017 54
December 2017 130
January 2018 158
February 2018 287
March 2018 306
April 2018 158
May 2018 150
June 2018 98
July 2018 82
August 2018 80
September 2018 394
October 2018 586
November 2018 228
December 2018 145
January 2019 508
February 2019 503
March 2019 568
April 2019 641
May 2019 218
June 2019 134
July 2019 135
August 2019 133
September 2019 176
October 2019 203
November 2019 171
December 2019 127
January 2020 144
February 2020 156
March 2020 217
April 2020 209
May 2020 87
June 2020 104
July 2020 62
August 2020 60
September 2020 91
October 2020 141
November 2020 112
December 2020 57
January 2021 93
February 2021 90
March 2021 126
April 2021 113
May 2021 63
June 2021 54
July 2021 44
August 2021 71
September 2021 181
October 2021 105
November 2021 78
December 2021 84
January 2022 80
February 2022 77
March 2022 56
April 2022 74
May 2022 81
June 2022 61
July 2022 46
August 2022 83
September 2022 91
October 2022 122
November 2022 94
December 2022 94
January 2023 84
February 2023 115
March 2023 96
April 2023 80
May 2023 62
June 2023 50
July 2023 24
August 2023 51
September 2023 66
October 2023 101
November 2023 66
December 2023 101
January 2024 71
February 2024 89
March 2024 115
April 2024 31

Citations

Powered by Dimensions

183 Web of Science

Altmetrics

×

Email alerts

Article activity alert

Advance article alerts

New issue alert

Receive exclusive offers and updates from Oxford Academic

Citing articles via

Google Scholar

  • Latest

  • Most Read

  • Most Cited

Redox regulation of epigenetic and epitranscriptomic gene regulatory pathways in plants
Sugar sensing in C4 source leaves: a gap that needs to be filled
Evolution of the FLOWERING LOCUS T-like genes in angiosperms: a core-Lamiales-specific diversification
Redox regulation of gene expression: proteomics reveals multiple previously undescribed redox-sensitive cysteines in transcription complexes and chromatin modifiers
OsRH52A, a DEAD-Box protein, is required for embryo sac development by regulating functional megaspore specification in rice

More from Oxford Academic

Biological Sciences

Plant Sciences and Forestry

Science and Mathematics

Books

Journals

Advertisem*nt

Plasticity in light reactions of photosynthesis for energy production and photoprotection (2024)

References

Top Articles
Latest Posts
Article information

Author: Jeremiah Abshire

Last Updated:

Views: 5883

Rating: 4.3 / 5 (54 voted)

Reviews: 85% of readers found this page helpful

Author information

Name: Jeremiah Abshire

Birthday: 1993-09-14

Address: Apt. 425 92748 Jannie Centers, Port Nikitaville, VT 82110

Phone: +8096210939894

Job: Lead Healthcare Manager

Hobby: Watching movies, Watching movies, Knapping, LARPing, Coffee roasting, Lacemaking, Gaming

Introduction: My name is Jeremiah Abshire, I am a outstanding, kind, clever, hilarious, curious, hilarious, outstanding person who loves writing and wants to share my knowledge and understanding with you.