Small Molecules Govern Thiol Redox Switches

Small Molecules Govern Thiol Redox Switches

TRPLSC 1699 No. of Pages 14 Opinion Small Molecules Govern Thiol Redox Switches Johannes Knuesting1 and Renate Scheibe1,* Oxygenic photosynthesis ga...

2MB Sizes 0 Downloads 69 Views

TRPLSC 1699 No. of Pages 14

Opinion

Small Molecules Govern Thiol Redox Switches Johannes Knuesting1 and Renate Scheibe1,* Oxygenic photosynthesis gave rise to a regulatory mechanism based on reversible redox-modifications of enzymes. In chloroplasts, such on–off switches separate metabolic pathways to avoid futile cycles. During illumination, the redox interconversions allow for rapidly and finely adjusting activation states of redox-regulated enzymes. Noncovalent effects by metabolites binding to these enzymes, here addressed as ‘small molecules’, affect the rates of reduction and oxidation. The chloroplast enzymes provide an example for a versatile regulatory principle where small molecules govern thiol switches to integrate redox state and metabolism for an appropriate response to environmental challenges. In general, this principle can be transferred to reactive thiols involved in redox signaling, oxidative stress responses, and in disease of all organisms.

Highlights Metabolic control is based on reversible redox-modifications leading to a dynamic steady state ratio of two enzyme forms differing in their activities. Fine-tuning of the activation state of each redox-controlled enzyme is achieved by non-covalently bound ‘small molecules’ which affect the rates of the redox-interconversions. ‘Small molecules’ such as substrates or products of the particular enzyme are able to shift redox potentials of the regulatory thiol groups by influencing the local microenvironment of these cysteine residues.

Redox Switches as the Basis for Cellular Homeostasis Redox chemistry is a central topic concerning all biological systems. In photoautotrophic organisms, during photosynthesis light energy is converted into strong reductants. At the same time, due to the asymmetrical architecture of the thylakoid membrane and vectorial electron flow, a proton motive force is generated which is used for ATP production. Reductants and ATP generated continuously during photosynthetic electron transport (PET) are needed for all cellular activities, in particular for biomass production by the C-, N-, and S-assimilatory processes. Since pool sizes of the involved reductants and energy carriers are extremely small, they need to be kept in a state that is ‘half full and half empty’ to allow for continuous fluxes under changing supply and demand. Therefore, flexible and fast regulation of key enzymes in all the energy-requiring pathways is essential to maintain metabolic and redox homeostasis. Light–dark modulation (see Glossary) of chloroplast enzymes by redox modifications at the various steps of the Calvin-Benson cycle (CBC) has been intensively investigated over the past decades. Originally it was thought to be a plant-specific type of regulation, to avoid a futile cycle with light and dark metabolism going on at the same time. But now, redox-controlled processes become apparent in all organisms. The historical development of the redox field over the past 50 years of research is provided in comprehensive reviews [1,2]. It started with the discovery of light–dark modulation in plants and its fine-tuning by ‘small molecules’. More recently, the redox field expanded also to non-plant research areas, finally leading to very active investigations in medical sciences where disturbance of redox homeostasis is causative for many diseases and aging [3,4]. Reversible thiol modifications are generally mediated by redox-active proteins, among them numerous members of the thioredoxin (Trx) superfamily that can transfer electrons to target proteins. Trx, in turn, mediate reduction and oxidation of redox-regulated target enzymes [5–7]. The chloroplasts of the model plant Arabidopsis thaliana contain a large number of Trx isoforms

Trends in Plant Science, Month Year, Vol. xx, No. yy

This principle of fine-tuning exemplified here for light-dark-modulated chloroplast enzymes is suggested for other proteins that are also subjected to reversible post-translational modifications.

1

Department of Plant Physiology, Faculty of Biology and Chemistry, Osnabrück University, Barbarastr. 11, 49076 Osnabrück, Germany

*Correspondence: [email protected] (R. Scheibe).

https://doi.org/10.1016/j.tplants.2018.06.007 © 2018 Elsevier Ltd. All rights reserved.

1

TRPLSC 1699 No. of Pages 14

as well as glutaredoxins (Grx) and peroxiredoxins [8,9]. Electrons for the reduction of the various Trx originate from the PET and are withdrawn at ferredoxin via ferredoxin-thioredoxin reductase (FTR), or from NADPH mediated by NADPH-Trx reductases NTRA (cytosol), NTRB (mitochondria), or NTRC (chloroplasts). The redox-regulated chloroplast enzymes are characterized by a high reactivity towards oxidants due to thermodynamic and kinetic effects that modify the properties of their thiols [10]. Furthermore, alkalization of the stroma of the photosynthesizing chloroplast increases deprotonation of the regulatory cysteines, leading to enhanced thiol reactivity. Therefore, continuous electron flow is required to maintain a certain portion of each enzyme in the reduced state. The chloroplast with the production of both highly reducing components, such as reduced ferredoxin (–420 mV), and oxygen (+816 mV) from photosynthetic water oxidation, appears to be perfectly suited to facilitate continuous thiol/disulfide cycling as a basis for a regulatory mechanism under nonstressed conditions with no increase in reactive oxygen species (ROS). Indeed, photosynthesis provides ample amounts of electrons, mostly in excess of ATP. The cost for regulation can be covered easily from light-generated electron flow, since the adjustment of the required ATP/NADPH ratio is achieved by various poising mechanisms [11,12]. In general, three types of thiol switches for regulation can be distinguished that exhibit different switching frequencies: (i) on–off switches between light and dark metabolism, (ii) continuous redox-cycling of redox-modulated chloroplast enzymes in the light at the expense of additional energy in the form of electrons, and (iii) redox sensors and signal transducers to prevent oxidative stress (Figure 1). In each case, the responsiveness of the switches can be influenced by small molecules that bind noncovalently. In this opinion paper, we present evidence from biochemical and physiological studies for the role of small molecules in rapidly and flexibly adjusting the redox switches. We will summarize the essential features of the chloroplast system, which is unique in some respect but offers general ideas to comprehend other regulatory systems. In the case of chloroplasts, the redox switches are not only operating as on–off switches (type 1) but provide the basis for fine-tuning (type 2). The effect of small molecules, originating from metabolic activities and environmental stress, on the redox interconversions allows for the integration of different types of information. Any disturbance of cellular homeostasis will result in changes of pool sizes of metabolites and inorganic ions. Change of the concentrations of these small molecules acts as indicators for the need to readjust metabolism to maintain homeostasis. Small molecules directly affect the interconversion rates of thiol switches and therewith the steady-state ratios of unmodified and modified protein, each of the two exhibiting different properties. Such a fine-controlled regulatory system is likely to also be realized in other cases where thiol switches have been found. To date, numerous further examples have become apparent, establishing the redox chemistry of thiol groups as the universal basis for controlling cellular homeostasis, sensing, and signal transduction [13–17].

Redox-Modulated Enzymes in Chloroplasts Light–dark-modulated chloroplast enzymes represent well-studied examples for redox regulation (Figure 2). Several enzymes of the CBC are activated by light-dependent reduction, mediated by Trx isoforms (mainly Trx m and f). Three of them catalyze practically irreversible reactions, namely fructose 1,6-bisphosphatase (FBPase; E.C.3.1.3.11), sedoheptulose 1,7bisphosphatase (SBPase; E.C.3.1.3.37), and phosphoribulose kinase (PRK; E.C.2.7.1.19), controlling fluxes through different parts of the branched pathway. The reductive phase of the CBC is catalyzed by a NAD(P)-dependent glyceraldehyde 3-phosphate dehydrogenase 2

Trends in Plant Science, Month Year, Vol. xx, No. yy

Glossary Calvin-Benson cycle (CBC): the main reductive process in chloroplasts for biomass production in the form of carbohydrates and carbon skeletons for all organic compounds by using light-generated NADPH and ATP. Carbon assimilation comprises three steps: (i) fixation of CO2, (ii) reduction of 3PGA for triose phosphate production, and (iii) regeneration of the CO2 acceptor (see also supplemental information online, Figure S1). Cysteine modifications: oxidation of Cys leads to intra- and intermolecular disulfide bridges, Sglutathionylation, S-nitrosylation, Ssulfhydration or results in sulfenic, sulfinic, or sulfonic acid residues, the latter one being irreversible. Cysteines are most reactive when deprotonated in alkaline (micro) environment. Light–dark modulation: the chloroplast-specific regulatory principle involving thiol-disulfide exchange post-translationally adjusts enzyme activities of metabolism during illumination and in darkness to avoid futile cycling of reductive (CBC) and oxidative pentose-phosphate (OPP) cycle. Malate valve: regeneration of the electron acceptor NADP+ is required to prevent overreduction of the photosynthetic electron transport chain and ROS formation. The key enzyme of the malate valve is NADPMDH, which uses NADPH to form malate that is transported out of the chloroplast, allowing for flexible adjustment of the NADPH/ATP ratio inside the chloroplast. This step needs to be strictly controlled because continuous NADPH supply is required, with the highest priority for 3-PGA reduction in the CBC (see also supplemental information online, Figure S2). Oxidative pentose-phosphate (OPP) pathway: glucose-6phosphate is oxidized generating NADPH, pentose phosphate, and CO2. In the illuminated chloroplast, this reaction would interfere with the CBC, resulting in futile cycling and ATP loss, if it would not be shut off in the light. Redox-active proteins: cysteines in proteins fulfill different roles for

TRPLSC 1699 No. of Pages 14

(A)

(B)

Diurnal cycle

Dark Light Dark Light

Connuous cycling

Light

(C)

Oxidave stress

Stress

Recovery

red.

PET

O2

PET

O2

PET ROS

NADPH

ox. Figure 1. Thiol switches Involved in Diverse Cellular Functions. Thiol switches operate in the context of light–dark cycles (A), during illumination (B), and upon oxidative stress (C). The turnover of the redox cycle of various thiol switches can be once a day (A), continuously during illumination (B), or exclusively in periods of stress (C). (A) Metabolism in illuminated and in darkened chloroplasts is separated by an on–off switch. Photosynthetic electron transport (PET) provides reductant. When electron flow stops upon darkening, regulatory thiol proteins are oxidized. This diurnal light–dark modulation prevents waste of energy. (B) During illumination, due to the continuous reoxidation of regulatory cysteines that are characterized by their high disposition to become oxidized, photosynthetic electrons are sacrificed to run the redox cycle continuously (see also supplemental information online, Figure S3). These particular thiol switches serve as the basis for fine-tuning activation states individually, depending on the metabolic situation at each regulated step. (C) Oxidative stress provokes disruption of cellular redox homeostasis and formation of reactive oxygen species (ROS). The transiently modified proteins can take over moonlighting functions and are involved in signaling and changing gene transcription. Upon reestablishment of redox homeostasis or for continued ROS scavenging and repair, NADPH reverses these modifications. ox., oxidized; red., reduced.

(NADP-GAPDH; E.C.1.2.1.13). RubisCO is indirectly redox-dependent due to the activity of RubisCO activase, which is a target of Trx. The F-ATP synthase (CF1; E.C.3.6.1.34) and NADPmalate dehydrogenase (NADP-MDH; E.C.1.1.1.82), both involved in poising the ATP-toNADPH ratio, are also activated by reduced Trx. The ATPase uses the proton gradient of the thylakoid membrane to generate ATP. The NADP-MDH of C3 plants as part of the malate valve [18] is involved in the export of surplus, reducing equivalents as malate from the chloroplast into the cytosol. All these enzymes are switched off by oxidation. Interestingly, oxidation of enzymes by Trx can also lead to activation, as is the case for glucose 6-phosphate dehydrogenase (G6PDH; E.C.1.1.1.49), the key enzyme of the oxidative pentose-phosphate (OPP) pathway. Light inactivation of the plastid isoforms of G6PDH ensures that the OPP pathway and CBC do not operate simultaneously, avoiding futile cycling. There are more Trx targets in plastids, such as ADP-glucose pyrophosphorylase (AGPase; E.C.2.7.7.27) to generate the activated substrate ADP-glucose for starch biosynthesis [19], and acetyl-CoA carboxylase (ACCase; E.C.6.4.1.2) for the synthesis of malonyl-CoA as the activated starter molecule to initiate fatty acid synthesis [20]. Furthermore, plastidial transcription and translation, and defense against oxidative stress, involve the other plastid isoforms Trx x, y, and z, as well as Trx-like proteins such as ACHT1 and 4 and CDSP32 in a redoxdependent manner or also redox-independently [21–26]. As another example, tetrapyrrole biosynthesis has been shown to be dependent upon NTRC and Trx at various steps [27]. Processes in other cellular compartments, such as nuclear transcription and cytosolic translation, also respond in many ways to redox-dependent signals [28,29], pointing to a broader

structure and function. On the one hand, they can build disulfide bridges to stabilize protein structures. These permanently oxidized proteins are found in endoplasmic reticulum, Golgi, vacuole, and apoplast. In general, the majority of the intracellular cysteine-containing proteins are permanently reduced. Upon oxidation they are irreversibly damaged. In contrast, some intracellular redox-active proteins undergo a reversible redox change, depending on the presence of reductants or oxidants. Reduced and oxidized proteins are characterized by different properties and the ratio of both forms determines the actual enzyme activity. Small molecules: thiol switches are affected by small molecules acting as positive or negative effectors of reduction or oxidation of the regulatory thiols of redox-active proteins. They comprise metabolites and ions that bind noncovalently to these proteins, thus linking the metabolic state and their redox interconversions. Thioredoxin (Trx): two cysteines in their catalytic center are common to members of this ubiquitous family of small soluble proteins. Each Trx can exist in the reduced dithiol and the oxidized disulfide state. The formation of the energetically favored 14-membered ring upon oxidation at the protruding active site makes Trx a perfect redox mediator. Trx systems: chloroplasts possess two systems for reduction of the Trx, namely ferredoxin-Trx reductase (FTR) and NADPH-Trx reductase (NTRC). Cytosol and mitochondria possess additional Trx and the NADPH-Trx reductases, NTRA and NTRB, respectively. Even the nucleus has been found to contain a Trx system. The plant redox network contains further redoxins, such as glutaredoxins (Grx) and peroxiredoxins (Prx), also mediating redox changes.

Trends in Plant Science, Month Year, Vol. xx, No. yy

3

TRPLSC 1699 No. of Pages 14

PET NADPH

7

ADP+Pi

OAA

ATP

NADP

NADPH

6

DicT

NADP

1

NADPGAPDH

2

FBPase

3

SBPase

4

PRK

5

RubisCO acvase

6

NADPMDH

7

ATPase

8

G6PDH

Malate CO2

5

3PGA 1,3bisPGA

Ru1,5bisP

NADPH

ADP ATP

4 3 Ru5P

CO2

NADP

1

S7P

Pi

Pi

SBP

TP TPT Pi

NADPH

6PGlu NADPH

8

2 G6P

F6P

Pi

FBP

Starch

Figure 2. Redox-Regulated Chloroplast Enzymes. The activities of the chloroplast enzymes shown in the scheme are regulated by redox interconversions. Some of the enzymes are reductively activated (yellow and orange), others are reductively inactivated (black). These reversible cysteine modifications are specifically suited to control metabolism and redox homeostasis in the daily light–dark cycle and under changing light regimes during the day. The Calvin-Benson cycle (CBC) as a branched, cyclic pathway is controlled at these essential steps in the different parts (yellow). The malate valve enzyme NADP-MDH, as well as the chloroplast ATP synthase, maintain the required ATP/NADPH ratio (orange). Oxidation leads to disulfide formation of the regulatory cysteines of these enzymes, resulting in their inactivation. In contrast, the key enzyme of the oxidative pentose phosphate (OPP) pathway, namely G6PDH (black), is inactive in the light, when reduced, and active in the dark, when its thiols are oxidized. DicT, dicarboxylate translocator; FBP, fructose1,6-bisphosphate; FBPase, fructose 1,6-bisphosphatase; G6PDH, glucose6-phosphate dehydrogenase; NADP-GAPDH, NADP-dependent glyceraldehyde3-phosphate dehydrogenase; NADP-MDH, NADP-malate dehydrogenase; SBPase, sedoheptulose1,7bisphosphatase;PET, photosynthetic electron transport; PGA, phosphoglycerate; PRK,phosphoribulose kinase; TPT, triosephosphate-phosphate translocator.

distribution of redox-controlled steps that are potentially also regulated by additional factors, such as small molecules, for flexible adjustment.

Dynamics of Reduction and Reoxidation as Influenced by Small Molecules Oxidation of the regulatory thiols occurs when light-dependent photosynthetic electron flow is shut down. Addition of oxygen to an anaerobic chloroplast suspension was shown to enhance oxidative inactivation upon darkening [7,30]. Due to the aerobic environment inside the chloroplast, the reduced cysteine residues of the redox-modified enzymes are readily reoxidized also during illumination. This applies even more under high light, when the flux through the CBC should be increased and not decreased in the presence of light-generated ROS, such as H2O2. Continuous re-reduction is required to maintain a certain portion of the enzyme in the reduced form under steady-state conditions in the light. Since specific metabolites act as positive or negative effectors on the reduction and oxidation rates of the interconversion in each case, the steady-state fraction of reduced enzyme is determined by the current metabolic

4

Trends in Plant Science, Month Year, Vol. xx, No. yy

TRPLSC 1699 No. of Pages 14

Net effect in steady state

NADP-GAPDH S S

HS

(A8B8)

+

(A8B8)

HS

1,3bisPGA + NADPH HS

(A2B2)

Feed-forward acvaon by 1,3bisPGA

HS GAP + NADP + Pi FBPase + FBP S

HS

S

HS

Feed-forward acvaon by FBP F6P + Pi

PRK -

Ru5P + ATP

S

HS

S

HS

Metabolismindependent acvaon Ru1,5bisP + ADP

NADP-MDH OAA + NADPH S

HS

S

HS

Feed-back inhibion of reducve acvaon by NADP

Malate + NADP

-

G6PDH + 6PGlu + HS

S

HS

S

NADPH

G6P + NADP

Feed-back acvaon of reducve inacvaon by NADPH

Figure 3. Redox-Regulated Chloroplast Enzymes Subjected to Individual Fine-Tuning by ‘Small Molecules’. The redox-regulated chloroplast enzymes shown in the schemes are reduced by the thioredoxin system in the light and continuously reoxidized due to the presence of oxygen. Small molecules (red), which are often substrates or products of the reaction catalyzed by the respective active enzyme (round symbol), modulate the rates of interconversion between (See figure legend on the bottom of the next page.) Trends in Plant Science, Month Year, Vol. xx, No. yy

5

TRPLSC 1699 No. of Pages 14

situation [31–33] (Figure 3). Additionally, as is true for most metabolic enzymes, catalysis of the active form itself is also influenced by substrates, products, cellular ion concentrations, and pH. The concept of fine-tuning an on–off switch, such as light–dark modulation of chloroplast enzymes, grew from work with the spinach chloroplast NADP-MDH. Interconversion of the active and inactive forms of NADP-MDH is directly influenced by the small molecule NADP+ [34]. Similarly, redox-dependent rates of activation of other chloroplast enzymes are controlled by metabolites [35]. Binding of a specific metabolite in each case was found to shift the oxidation constant (Kox) for the respective regulatory cysteines after equilibration with dithiothreitol (DTT)-redox buffers [36]. Such experiments, as well as kinetic studies, have provided the first evidence that a common principle (redox-cycling) is used to control fluxes via small molecules, specifically at defined metabolic steps, according to the demand [37]. Effectordriven change in the steady-state concentration of the active form of the enzyme allows for rapid adjustment of fluxes. Local metabolite concentrations are direct indicators of the actual carbon flux at each of the redox-regulated steps. Functioning as positive or negative effectors, they immediately can readjust the fluxes specifically upon any imbalance in metabolite pools, in this way maintaining homeostasis. The redox-dependent changes of the kinetic parameters can concern Vmax (NADP-MDH, PRK), Km (NADP-GAPDH, FBPase, G6PDH, and AGPase), or Ki (AGPase, RubisCO activase). The monocascade of the continuous redox cycle between the two enzyme forms enables a hysteretic response within less than a minute to gradually or rapidly alter actual enzyme activities as required (Figure 3). NADP-GAPDH The chloroplast NADP-GAPDH catalyzes the reduction of 3-PGA, which is achieved after its conversion into 1,3-bisphosphoglycerate (1,3bisPGA) by 3-phosphoglycerate kinase (PGK). The active enzyme consists of two different subunits, GapA and GapB, in a heterotetramer (A2B2). The dark form with two regulatory thiols in the oxidized state exists as a hexadecameric complex (A8B8), which is virtually inactive under physiological substrate (1,3bisPGA) concentrations. Only the simultaneous presence of reduced Trx and 1,3bisPGA, acting as an effector, leads to the dissociated tetrameric form with an increased affinity for the substrate 1,3bisPGA [38,39]. For GapA/B, reduction of the disulfide bridge at the C terminal extension of GapB in the hexadecameric complex lowers the activation constant (Ka) for 1,3bisPGA, acting here as a positive effector to facilitate dissociation and activation at physiological 1,3bisPGA concentrations. In addition, another complex consisting of inactive NADP-GAPDH, PRK, and the small chloroplast protein CP12 is even more readily activated by reduction and dissociation in the presence of NADPH already in low light. This leads to activation of a fraction of the total NADP-GAPDH pool and all of the PRK, even in low light [40]. In contrast, the A8B8 complex releases the active enzyme only when the substrate 1,3bisPGA starts to accumulate, indicating oxidized and reduced forms. Redox cycling in the light, when reduction and concomitant reoxidation lead to a steady state of reduced and oxidized forms, provides the basis for individual fine-tuning by metabolites. Small molecules act as effectors (positive or negative) on either reduction or oxidation or both. The equilibrium between active and inactive enzyme can be shifted according to the nature of the effectors, as indicated by the arrows and the + and affecting the rates of the interconversions. The net effect is that a very rapid and efficient adjustment of enzyme activities is achieved according to the actual metabolic demand. This can result in feed-forward activation when substrate pools increase (NADP-GAPDH, FBPase), or feedback inhibition from the increase of product (NADP-MDH or G6PDH). Oxidation in the dark leads to disulfide formation of the regulatory cysteine residues. In most cases (NADP-GAPDH, FBPase, PRK, NADP-MDH), the oxidized enzymes are inactive (square symbols), except for G6PDH which is active upon oxidation. FBP, fructose 1,6bisphosphate; FBPase, fructose 1,6-bisphosphatase; F6P, fructose 6-phosphate; GAP, glyceraldehyde 3-phosphate; G6P, glucose 6-phosphate; G6PDH, glucose 6-phosphate dehydrogenase; NADP-GAPDH, NADP-dependent glyceraldehyde 3-phosphate dehydrogenase; NADP-MDH, NADP-malate dehydrogenase; OAA, oxaloacetate; PGA, 3-phosphoglycerate; PRK, phosphoribulose kinase; 1,3bisPGA, 1,3-bisphosphoglycerate; Ru1,5bisP, ribulose 1,5bisphosphate; Ru5P, ribulose 5-phosphate; 6PGlu, 6-phosphogluconate.

6

Trends in Plant Science, Month Year, Vol. xx, No. yy

TRPLSC 1699 No. of Pages 14

the need for an increased flux through this step. Interestingly, this results in NADP-GAPDH activities that correspond to the actual flux rate of the whole CBC [41]. In this context, CP12 itself can be seen as a specific redox regulator. Due to its properties as an intrinsically unstructured protein, carrying two disulfide bridges, it assists the oligomerization of a specific NADP-GAPDH/PRK complex upon oxidation and releases both CBC enzymes in their active state upon the onset of illumination [42–44]. PRK PRK is a unique enzyme in chloroplasts for regeneration of the CO2 acceptor in the CBC. It is rapidly activated in the light upon release from the NADP-GAPDH/CP12/PRK complex [40]. Both reductive activation and oxidative inactivation of PRK in vitro, using the purified dimeric form, is slowed down in the presence of ATP [45]. This results in no net effect of ATP on the activation state of PRK, which is fully reduced and active in standard assay medium already in low light [40]. Actual fluxes through this step are determined by metabolic inhibitors (inorganic phosphate, FBP, 6-phosphogluconate) at the level of catalysis of the fully reduced enzyme [46]. FBPase Chloroplast FBPase is reductively activated, but only when its substrate FBP is present as an effector [47]. In addition, the rate of oxidative inactivation is decreased in the presence of FBP [48]. FBP binding impacts thiol sensitivity of the distant vicinal thiol groups by long distance conformational changes, resulting in facilitated reduction and the stabilization of the reduced form of FBPase [49]. This results in an increased activation state of FBPase as soon as substrate starts to accumulate. Reduced FBPase is characterized by increased affinities for FBP and Mg2+, with [Mg2+-FBP]2 being the true substrate of the reaction, and by a pH optimum for catalysis in a more physiological range [50]. These shifts of Km (FBP; Mg2+) and pH optimum, together with the facilitated reduction, result in an increased FBPase activity in illuminated chloroplasts. NADP-MDH Poising mechanisms are required when the ATP demand in the chloroplast exceeds its production and NADPH accumulates. Under this condition, NADPH is reoxidized by NADP-MDH and malate is formed, resulting in indirect export of reducing equivalents as malate. In the cytosol, malate can be reoxidized, leading to the formation of NADH for nitrate reduction, or, in the mitochondria, for ATP production [18]. Reductive activation of NADP-MDH is only enabled when NADPH starts to accumulate in the stroma (e.g., due to the lack of ATP or CO2 for continued CBC activity), and when NADP+ levels decrease correspondingly. In the lag phase of photosynthesis, upon a switch from dark to light, overreduction of the electron transport chain, monitored as fluorescence emission from photosystem II, is counteracted by the fast activation of NADP-MDH and typical oscillations that are dampened in the next few minutes to reach a new steady-state [51]. NADP+ acts as a negative effector for light activation of NADP-MDH, indicating the preferential consumption of reducing equivalents in the CBC or by competing electron acceptors such as nitrite [52,53]. This is an essential feature of the selfcontrolled malate valve. NADPH in contrast, inhibits both activation as well as inactivation, thus decreasing the frequency of the redox cycle, but without any net effect on the ratio between reduced and oxidized form. Upon decrease of light intensity from high to lower light, a short peak of NADP+ accumulation and a transient decrease of O2 evolution coincides with a rapid decrease of NADP-MDH activity and the re-establishment of a new steady-state activity [54]. The oxidized enzyme is characterized by a Vmax of zero, where the oxidized C terminus of NADP-MDH is acting as an autoinhibitory peptide [55]. The C terminus is released from the active site upon reduction only in the absence of the negative effector NADP+ [56].

Trends in Plant Science, Month Year, Vol. xx, No. yy

7

TRPLSC 1699 No. of Pages 14

G6PDH Reductive inactivation of chloroplast G6PDH rapidly and efficiently occurs when NADPH is sufficiently provided in the light reactions; here NADPH acts as a positive effector of the reductive inactivation. The reduced inactive G6PDH is characterized by its low affinity towards the substrate G6P (Km > 20 mM), rendering it practically inactive at physiological G6P concentrations. It is worth mentioning that this enzyme is active when its regulatory thiols are oxidized by the formation of a disulfide bridge; a unique case for an intracellular enzyme. Here, oxidation results in a shift of the Km value to the physiological range (1–2 mM) of the substrate G6P [57,58]. Other Enzymes Redox state and small molecules appear to act together in other cases also. The carboxylation step in the CBC, catalyzed by RubisCO, is regulated by the ATP-dependent RubisCO activase, which removes incorrect metabolites from the active site [59]. The reduced form of the activase is less affected by the inhibitor ADP. Therefore, the carboxylation activity can be adjusted to the conditions as indicated by the ADP/ATP ratio, particularly in low light [60]. The g-subunit of chloroplast CF0/CF1-ATP synthase undergoes a redox change of two cysteine residues, positioned in a chloroplast-specific amino acid sequence insertion, thus preventing ATP hydrolysis in the dark [61–63]. However, ATP synthesis is not only redox controlled, but also responds to pH gradient, proton motive force, and other metabolic factors. Ambient conditions, such as low light or elevated CO2, can therefore influence the ease of reductive activation and thus the rate of ATP synthesis [64,65]. The reductive activation of AGPase is promoted by the regulatory metabolite trehalose 6-phosphate (T6P) [66]. For glutamine synthase (GS), glutamate synthetase (GOGAT), and acetyl-CoA carboxylase (ACCase), similar mechanisms of an integrated activation have been suggested but remain to be studied in more detail. Indeed, it is experimentally challenging to identify the site of action of the effectors and their role in shifting the equilibrium redox potential, particularly when activation and activity cannot be tested separately. It is also difficult to distinguish and characterize the different mechanisms of regulation, when substrate or products act simultaneously as effectors on the redox interconversion or when both catalysis and activation/inactivation are influenced by the same small molecules. Taken together, small molecules act as positive or negative effectors on reduction and/or oxidation rates of the redox interconversions. This principle allows for the individual adjustment of the redox states and concomitantly the properties of the respective proteins (Figure 4A, Key Figure). As an example, for the CBC enzyme FBPase and the malate-valve enzyme NADPMDH, this can result in a diverging effect concerning their activation state and consequently their actual activities (Figure 4B). The extraordinarily negative redox potential of reduced ferredoxin is capable of reducing these enzymes, however, only in the presence of the positive effector (FBP for FBPase) and the absence of the negative effector (NADP+ for NADP-MDH). Photosynthetic thylakoid energization and redox state of the electron transport chain are thus tightly interconnected with the redox cycles of target enzymes. The control of these monocascades by metabolic intermediates results in a continuous adjustment and maintenance of fluxes under challenging conditions.

Concluding Remarks and Future Perspectives Usually in lab experiments, when huge effects are anticipated for clear-cut results, a large impact is applied to produce visible displacements which provoke system disturbance, so that oscillations and subsequent dampening can be easily observed [51]. In reality, however, infinitely small, immediate changes of the activation states result in a smooth adjustment of 8

Trends in Plant Science, Month Year, Vol. xx, No. yy

TRPLSC 1699 No. of Pages 14

Key Figure

Opposite Effects on Calvin-Benson Cycle (CBC) and Malate Valve Activation Depending on the Metabolic State (A)

Small molecules (posive or negave effector) 2e-+ 2H+

S

SH

S

SH

Small molecules (posive or negave effector)

H2O

½ O2

(B)

-420 NADP-MDH

FBPase

Fdox/Fdred

S

S

-400

[NADP+]

S S

mV

DTTox/ -380 DTT red Trxm

-360

S

Trxf

SH

[FBP]

-340

SH

S

S S

SH SH

NAD(P)/ -320 NAD(P)H

-300 -250 GSSG/GSH -200

(See figure legend on the bottom of the next page.) Trends in Plant Science, Month Year, Vol. xx, No. yy

9

TRPLSC 1699 No. of Pages 14

the steady-state activities and fluxes, which are tightly linked to the metabolic situation. The complexity of the regulatory network that controls energy distribution, biomass production, antioxidant defense, gene transcription, and translation relies on the presence of many redox transmitters. The interplay between redox systems and metabolism is also the basis of genetic and epigenetic mechanisms in the context of adaptive responses. At this level, but also in the classic metabolic cases of redox-based modifications as described in this article, the combinatorial aspect of post-translational modifications and allosteric effects with gradual impact on each other might explain the robustness of plant metabolism [67]. While the plastid Trx are characterized by exceptionally negative redox potentials in the range of 330 to 370 mV (at stromal pH 7.9) [68], cytosolic Trx and all Grx range between 220 and 270 mV (at pH 7) [69]. Redox potentials of the cysteines in the C-X-X-C motif of the Trx-fold can vary dynamically in the range between 95 mV (protein-disulfide isomerase in the endoplasmatic reticulum) and 287 mV (cytTrx) [70], or even be more negative as in chloroplast Trx f and m. As one pH unit changes the redox potential of cysteines by 60 mV [71], values from measurements performed at pH 7, as done in many studies, should be corrected to a more negative value when compared with the values of the plastid isoforms, namely to 280 to 330 mV. It is interesting to note that the midpoint potentials of chloroplast Trx with their rather negative values can only be determined in redox buffers with similar oxidation constants (Kox). In fact, only DTT-redox buffers are suitable for this purpose. Finally, in a steady-state system, red/ ox ratios are not only determined by their equilibrium redox potentials (by redox titrations), but also by the kinetics of reduction and oxidation [72]. It is still a challenge to assign specific targets to each of the many Trx. Some specificity of Trx might be explained by differences in their redox potentials, those of the various plastidial Trx isoforms ranging around 360 mV for Trx m and f [68] and below 310 mV for Trx x, y, and z [73]. However, the main driving force for target recognition appears to lie in the surface properties that facilitate the noncovalent interaction between Trx and the target [74,75]. Genetic studies and in vitro experiments using the recombinant proteins indicate some specificity of plastidial Trx isoforms for different targets [73,76]. Furthermore, mutant plants with a changed expression of the various Trx systems showed deviations in photosynthetic performance, Figure 4. (A) The redox cycle of protein dithiol-disulfides is driven by electron flow from reductants to an oxidant. Such a seemingly futile cycle serves as a basis for fine-tuning. The steady-state concentration of active enzyme can be readily shifted by ‘small molecules’ that either positively or negatively act as effectors on the rates of interconversion. (B) Chloroplast components are arranged according to their midpoint potentials. The disulfide bridges of light–dark-regulated enzymes are reduced by the thioredoxins Trx m and f, which obtain electrons from the strong reductant Fd (ferredoxin). Differential activation of FBPase (fructose 1,6-bisphosphatase) and NADP-MDH (NADP-malate dehydrogenase) are shown as examples for individual fine-tuning. Changing stromal concentrations of FBP (fructose 1,6-phosphate) and NADP+ are visualized by color gradients. In the case of FBPase, a low concentration of the positive effector FBP leads to a very negative redox potential of the regulatory cysteines and hinders enzyme activation by thioredoxin (Trx). FBP (blue oval) at increasing concentrations binds to the active site and, by an allosteric effect (marked by the grey arrow), shifts the redox potential of the regulatory cysteines to a less negative value, so that FBPase can be easily reduced. An increasing FBP concentration represents the carbohydrate pool to be further processed in the chloroplast and opens the gate by facilitating FBPase activation. When the flux through the CBC is high and light-generated NADPH is continuously used for CO2 assimilation, NADP+ (red oval) as a negative effector impedes reductive activation of NADP-MDH and the autoinhibitory C terminus blocks the active site, rendering the enzyme inactive. As a consequence, export of NADPH as malate is impeded when reductants are preferentially used in the CBC, and increased NADP acts as a brake to avoid inappropriate opening of the malate valve by inhibition of reductive activation. In contrast, a low NADP+/(NADP+ + NADPH) ratio (corresponding to high NADPH), renders the C terminal disulfide bridge accessible and shifts the redox potential to a less negative value, thus enabling reductive activation of NADP-MDH. DTT, dithiothreitol; FBP, fructose 1,6-bisphosphate; GSSG/GSH, oxidized and reduced glutathione; ox, oxidized; red, reduced.

10

Trends in Plant Science, Month Year, Vol. xx, No. yy

TRPLSC 1699 No. of Pages 14

enzyme activation kinetics, and adaptation to fluctuating light conditions, but are also suggesting partial redundancy [77–79]. In an in vitro system, FTR and NTRC reduce the various chloroplast Trx isoforms more or less efficiently [80]. However, the well-studied chloroplast redox systems and their possible targets, particularly their specificities for the targets, are not yet fully understood [81–83]. The availability of crystal structures of redox-regulated targets improves the knowledge of the molecular details of these proteins. Their sequence characteristics and special properties came into existence during the period when oxygenic photosynthesis created conditions that were favorable for using oxidation and re-reduction as a means to post-translationally change enzyme properties [84]. During evolution, upon the emergence of oxygenic photosynthesis, some chloroplast enzymes acquired sequence extensions or inserts carrying the regulatory thiols, while the remaining sequence is highly similar to the counterparts that are not subjected to redox modulation [84,85]. The extra sequences carry two vicinal thiols and act as a redox switch. Upon reversible modification, local restriction or a conformational impact changes the catalytic or regulatory properties of the enzyme (Figure 4B). Some structures of the respective enzymes in their reduced and oxidized forms are already available, like those for FBPase [49,86] and for NADP-MDH [87,88] with NADP+ bound to the oxidized form in such a way that disulfide reduction is impeded [6]. From crystal structures of two very similar Trx h isoforms from barley, the importance of hydrogen bonds between backbones of Trx and target protein has been suggested as a determinant for specific recognition [89,90]. Moreover, the various Trx h isotypes were also shown to possess specific recognition sites for their targets [91]. To explain the specificity among the unusually high number of redox systems and their targets, the in vivo situation with its high protein concentration promoting protein–protein interactions should be taken into account. More specific protein–protein interactions might, therefore, be achieved only in vivo. Steric constraints for disulfide formation can be increased or decreased by noncovalent interactions that are affected by ligand binding or additional post-translational modifications (PTM code) [92–95], thus changing the reductive and oxidative properties, respectively, of each redox mediator protein. In order to predict a change of redox potential, the interplay of bound small molecules, additional modifications, and other factors should be taken into account. All these factors can shift the pKa of the respective thiols. Also, steric constraints can impede disulfide formation, and noncovalent binding of effectors influences the local electrostatic environment and affects thiolate formation [96–100]. Binding of small molecules puts further impact on the reactivity of the thiol(s), as shown for OOC-5, a Torsin-family protein in Caenorhabditis elegans, where ADPbinding shifts the redox potential from 210 to 240 mV. In this way, by integrating the ADP/ ATP ratio and redox state, this protein acts as a sensor for metabolic requirements [101]. The regulatory principle described here for well-studied examples is not only present in chloroplasts of plants growing under challenging environmental conditions, which require the continuous and rapid adjustment of enzyme activities by post-translational modification. Similar redox-regulatory mechanisms also exist in yeast and animals during oxidative stress affecting many cellular functions, including signaling and regulation of nuclear transcription [70,102–105]. Redox changes of intracellular proteins occur in all organisms, particularly in the context of development [106], aging, pathogen attack, and disease [107]. Better knowledge of the ubiquitous redox switches and their appropriate adjustment in complex regulatory Trends in Plant Science, Month Year, Vol. xx, No. yy

11

TRPLSC 1699 No. of Pages 14

networks will enable advances in all fields, including medical sciences (see Outstanding Questions). Acknowledgments Work performed in the author’s lab was supported with funds from the DFG (SCHE 217; SPP 1710) over the past four decades.

Supplemental Information Supplemental information associated with this article can be found, in the online version, at https://doi.org/10.1016/j. tplants.2018.06.007.

References 1.

Buchanan, B.B. et al. (2012) Fifty years in the thioredoxin field and a bountiful harvest. Biochim. Biophys. Acta 1820, 1822–1829

2.

Buchanan, B.B. (2016) The path to thioredoxin and redox regulation in chloroplasts. Annu. Rev. Plant Biol. 67, 1–24

21. Broin, M. et al. (2002) The plastidic 2-cysteine peroxiredoxin is a target for a thioredoxin involved in the protection of the photosynthetic apparatus against oxidative damage. Plant Cell 14, 1417–1432

3.

Go, Y.M. and Jones, D.P. (2017) Redox theory of aging: implications for health and disease. Clin. Sci. (Lond.) 131, 1669–1688

22. Danon, A. (2002) Redox reactions of regulatory proteins: do kinetics promote specificity? Trends Biochem. Sci. 24, 197–203

4.

McBean, G.J. et al. (2017) Redox-based therapeutics in neurodegenerative disease. Br. J. Pharmacol. 174, 1750–1770

5.

Buchanan, B.B. and Balmer, Y. (2005) Redox regulation: a broadening horizon. Annu. Rev. Plant Biol. 56, 187–220

23. Dangoor, I. et al. (2012) A chloroplast light-regulated oxidative sensor for moderate light intensity in Arabidopsis. Plant Cell 24, 1894–1906

6.

Dai, S. et al. (2000) How does light regulate chloroplast enzymes? Structure-function studies of the ferredoxin/thioredoxin system. Q. Rev. Biophys. 33, 67–108

7.

Schürmann, P. and Buchanan, B.B. (2008) The ferredoxin/thioredoxin system of oxygenic photosynthesis. Antioxid. Redox Signal. 10, 1–39

8.

Meyer, Y. et al. (2008) Glutaredoxins and thioredoxins in plants. Biochim. Biophys. Acta 1783, 589–600

9.

Ströher, E. and Millar, A.H. (2012) The biological roles of glutaredoxins. Biochem. J. 446, 333–348

10. Gilbert, H.F. (1990) Molecular and cellular aspects of thiol-disulfide exchange. In Advances in Enzymology (Meister, A., ed), Vol. 63, pp. 69–172, John Wiley & Sons 11. Alric, J. and Johnson, X. (2017) Alternative electron transport pathways in photosynthesis: a confluence of regulation. Curr. Opin. Plant Biol. 37, 78–86 12. Hanke, G. and Scheibe, R. (2018) The contribution of electron transfer after photosystem I to balancing photosynthesis. In Photosynthesis and Bioenergetics (Barber, J. and Ruban, A. V., eds), pp. 277–303, World Scientific Publishing 13. Waszczak, C. et al. (2018) Reactive oxygen species in plant signaling. Annu. Rev. Plant Biol. 69, 209–236 14. Cremers, C.M. and Jakob, U. (2013) Oxidant sensing by reversible disulfide bond formation. J. Biol. Chem. 288, 26489–28496 15. Noctor, G. et al. (2017) ROS-related redox regulation and signaling in plants. Sem. Cell. Dev. Biol. 80, 3–12 16. Lillig, C.H. and Berndt, C. (2013) Glutaredoxins in thiol/disulfide exchange. Antioxid. Redox Signal. 18, 1654–1665 17. Mullineaux, P.M. et al. (2018) ROS-dependent signaling pathways in plants and algae exposed to high light: comparisons with other eukaryotes. Free Radic. Biol. Med. Published online 2, 2019. http://dx.doi.org/10.1016/j. February freeradbiomed.2018.01.033 18. Scheibe, R. (2004) Malate valves to balance cellular energy supply. Physiol. Plant 120, 21–26 19. Michalska, J. et al. (2009) NTRC links built-in thioredoxin to light an sucrose in regulating starch synthesis in chloroplasts and amyloplasts. Proc. Natl. Acad. Sci. U. S. A. 106, 9908–9913 20. Kozaki, A. et al. (2001) Thiol-disulfide exchange between nuclear-encoded and chloroplast-encoded subunits of pea acetyl-CoA carboxylase. J. Biol. Chem. 276, 39919–39925

12

Trends in Plant Science, Month Year, Vol. xx, No. yy

24. Eliyahu, E. et al. (2015) ACHT4-driven oxidation of APS1 attenuates starch synthesis under low light intensity in Arabidopsis plants. Proc. Natl. Acad. Sci. U. S. A. 112, 12876–12881 25. Dietz, K.J. and Pfannschmidt, T. (2011) Novel regulators in photosynthetic redox control of plant metabolism and gene expression. Plant Physiol. 155, 1477–1485 26. Serrato, A.J. et al. (2013) Plastid thioredoxins: a “one-for-all” redox-signaling system in plants. Front. Plant Sci. 4, 463 27. Da, Q. et al. (2017) Thioredoxin and NADPH-dependent thioredoxin reductase C regulation of tetrapyrrole biosynthesis. Plant Physiol. 175, 652–666 28. Moore, M. et al. (2015) Redox regulation of cytosolic translation in plants. Trends Plant Sci. 21, 388–397 29. Shaikhali, J. and Wingsle, G. (2017) Redox-regulated transcription in plants: emerging concepts. AIMS Mol. Sci. 4, 301–338 30. Leegood, R.C. and Walker, D.A. (1980) Regulation of fructose1,6-biphosphatase activity in intact chloroplasts. Biochim. Biophys. Acta 593, 362–370 31. Scheibe, R. et al. (1989) Chloroplast glucose-6-phosphate dehydrogenase: Km shift upon light modulation and reduction. Arch. Biochem. Biophys. 274, 290–297 32. Holtgrefe, S. et al. (1997) Regulation of steady-state photosynthesis in isolated intact chloroplasts under constant light: responses of carbon fluxes, metabolic pools and enzyme-activation states to changes of electron pressure. Plant Cell Physiol. 38, 1207–1216 33. Scheibe, R. (1990) Redox-regulation of chloroplast enzymes: mechanism and physiological significance. In Current Research in Photosynthesis (Baltscheffsky, M., ed), Vol. 3, pp. 127–134, Kluwer Academic Publishers 34. Scheibe, R. and Jacquot, J.P. (1983) NADP-regulates the light activation of NADP-dependent malate dehydrogenase. Planta 157, 548–553 35. Scheibe, R. (1995) Light modulation of stromal enzymes. In Carbon Partitioning and Source-Sink Interactions in Plants (Madore, M.A. and Lucas, W.J., eds), pp. 13–22, The American Society of Plant Physiologists 36. Faske, M. et al. (1995) Redox equilibria between the regulatory thiols of light/dark-modulated chloroplast enzymes and dithiothreitol: fine-tuning by metabolites. Biochim. Biophys. Acta 1247, 135–142 37. Scheibe, R. (1991) Redox-modulation of chloroplast enzymes: a common principle for individual control. Plant Physiol. 96, 1–3

Outstanding Questions Are there additional thiol-switches in signaling, subcellular localization or gene expression that are controlled by ‘small molecules’? Does local influence on thiolate formation from further yet unknown modifications of neighboring amino acids need to be taken into account to fully understand the dynamic behavior of a highly integrated regulatory network? Is the principle of fine-tuning by ‘small molecules’ also valid in other cases of post-translationally regulated proteins, for example, nitrate reductase, sucrose-P synthase and PEP-carboxylase which are subject to protein phosphorylation/-dephosphorylation? Is specificity in vivo reached by local concentrations of ‘small molecules’ such as protons, inorganic ions, metabolites, or further covalent modifications, or protein-protein interactions? Which other sterical and/or biochemical features determine the observed shifts of the redox potentials of the regulatory cysteines? Can simulations and modeling of the respective transient holo-complexes help to better understand and predict specificities, new targets, and functions?

TRPLSC 1699 No. of Pages 14

38. Baalmann, E. et al. (1995) Reductive modification and nonreductive activation of purified spinach chloroplast NADP-dependent glyceraldehyde-3-phosphate dehydrogenase. Arch. Biochem. Biophys. 324, 201–208 39. Baalmann, E. et al. (1994) Regulation of NADP-dependent glyceraldehyde 3-phosphate dehydrogenase activity in spinach chloroplasts. Bot. Acta 107, 313–320 40. Scheibe, R. et al. (2002) Co-existence of two regulatory NADPglyceraldehyde 3-P dehydrogenase complexes in higher plant chloroplasts. Eur. J. Biochem. 269, 5617–5624 41. Fridlyand, L.E. et al. (1997) Quantitative evaluation of the rate of 3-phosphoglycerate reduction in chloroplasts. Plant Cell Physiol. 38, 1177–1186 42. Gontero, B. and Avilan, L. (2011) Creating order out of disorder: structural imprint of GAPDH on CP12. Structure 19, 1728–1729 43. Erales, J. et al. (2009) CP12 from Chlamydomonas reinhardtii, a permanent specific “chaperone-like” protein of glyceraldehyde-3phosphate dehydrogenase. J. Biol. Chem. 284, 12735–12744 44. Trost, P. et al. (2006) Thioredoxin-dependent regulation of photosynthetic glyceraldehyde-3-phosphate dehydrogenase: autonomous vs. CP12-dependent mechanisms. Photosynth. Res. 89, 263–275 45. Ashton, A.R. (1975) Ribulose-5-phosphate kinase from Zea mays. Regulation of enzyme activity by oxidation and reduction. In Thioredoxins: Structure and Functions (Gadal, P., ed.), pp. 245–250, Editions du CNRS 46. Gardemann, A. et al. (1983) Regulation of spinach ribulose-5phosphate kinase by stromal metabolite levels. Biochim. Biophys. Acta 722, 51–60 47. Hertig, C.M. and Wolosiuk, R.A. (1983) Studies on the hysteretic properties of chloroplast fructose-1,6-bisphosphatase. J. Biol. Chem. 258, 984–989 48. Clancey, C.J.G. and Gilbert, H.F. (1987) Thiol/disulfide exchange in the thioredoxin-catalyzed reductive activation of spinach chloroplast fructose-1,6-bisphosphatase. J. Biol. Chem. 262, 13545–13549

modifying substrate accessibility and cofactor binding. Biochem. J. 457, 117–125 59. Zhang, N. and Portis, A.R., Jr (1999) Mechanism of light regulation of Rubisco: a specific role for the larger Rubisco activase isoform involving reductive activation by thioredoxinf. Proc. Natl. Acad. Sci. U. S. A. 96, 9438–9443 60. Zhang, N. et al. (2002) Light modulation of Rubisco in Arabidopsis requires a capacity for redox regulation of the larger Rubisco activase isoform. Proc. Natl. Acad. Sci. U. S. A. 99, 3330–3334 61. Wu, G. et al. (2007) A point mutation in atpC1 raises the redoxpotential of the Arabidopsis chloroplast ATP synthase g-subunit regulatory disulfide above the range of thioredoxin modulation. J. Biol. Chem. 282, 36782–36789 62. Hisabori, T. et al. (2013) The chloroplast ATP synthase features the characteristic redox regulation machinery. Antioxid. Redox Signal. 19, 1846–1854 63. Buchert, F. et al. (2015) Redox regulation of CF1-ATPase involves interplay between the gamma-subunit neck region and the turn region of the betaDELSEED-loop. Biochim. Biophys. Acta 1847, 441–450 64. Kohzuma, K. et al. (2013) Light- and metabolism-related regulation of the chloroplast ATP synthase have distinct mechanisms and functions. J. Biol. Chem. 288, 13156–13163 65. Carrillo, L.R. et al. (2016) Multi-level regulation of the chloroplast ATP synthase: the chloroplast NADPH thioredoxin reductase C (NTRC) is required for redox modulation specifically under low irradiance. Plant J. 87, 654–663 66. Kolbe, A. et al. (2005) Trehalose 6-phosphate regulates starch synthesis via posttranslational redox activation of ADP-glucose pyrophosphorylase. Proc. Natl. Acad. Sci. U. S. A. 102, 11118–11123 67. Lothrop, A.P. et al. (2013) Deciphering post-translational modification codes. FEBS Lett. 587, 1247–1257 68. Collin, V. et al. (2003) The Arabidopsis plastidial thioredoxins: new functions and new insights into specificity. J. Biol. Chem. 278, 23747–23752

49. Gütle, D.D. et al. (2016) Chloroplast FBPase and SBPase are thioredoxin-linked enzymes with similar architecture but different evolutionary histories. Proc. Natl. Acad. Sci. U. S. A. 113, 6779–6784

69. Bréhelin, C. et al. (2004) Cytosolic, mitochondrial thioredoxins and thioredoxin reductases in Arabidopsis thaliana. Photosynth. Res. 79, 295–304

50. Minot, R. et al. (1982) The role of pH and magnesium concentration in the light activation of chloroplastic fructose bisphosphatase. FEBS Lett. 142, 118–120

70. Wouters, M.A. et al. (2010) Disulfides as redox switches: from molecular mechanisms to functional significance. Antioxid. Redox Signal. 12, 53–91

51. Scheibe, R. and Stitt, M. (1988) Comparison of NADP-malate dehydrogenase activation, QA reduction and O2 evolution in spinach leaves. Plant Physiol. Biochem. 26, 473–481

71. Hirasawa, M. et al. (1999) Oxidation-reduction properties of chloroplast thioredoxins, ferredoxin:thioredoxin reductase, and thioredoxin f-regulated enzymes. Biochemistry 38, 5200–5205

52. Backhausen, J.E. et al. (2000) Electron acceptors in isolated intact spinach chloroplasts act hierarchically to prevent overreduction and competition for electrons. Photosynth. Res. 64, 1–13 53. Backhausen, J.E. et al. (1994) Competition between electron acceptors in photosynthesis: regulation of the malate valve during CO2 fixation and nitrite reduction. Photosynth. Res. 42, 75–86 54. Stitt, M. et al. (1989) Response of photosynthetic electron transport and carbon metabolism to a sudden decrease of irradiance in the saturating or the limiting range. Biochim. Biophys. Acta 973, 241–249 55. Miginiac-Maslow, M. et al. (1985) Energetic aspects of the light activation of two chloroplast enzymes: fructose-1,6-bisphosphatase and NADP-malate dehydrogenase. Photosynth. Res. 6, 201–213 56. Schepens, I. et al. (2000) Inhibition of the thioredoxin-dependent activation of the NADP-malate dehydrogenase and cofactor specificity. J. Biol. Chem. 275, 20996–21001 57. Wenderoth, I. et al. (1997) Identification of the cysteine residues involved in redox modification of plant plastidic glucose-6-phosphate dehydrogenase. J. Biol. Chem. 272, 26985–26990 58. Née, G. et al. (2014) Redox regulation of chloroplastic G6PDH activity by thioredoxin occurs through structural changes

72. Kemp, M. et al. (2008) Nonequilibrium thermotynamics of thiol/ disulfide redox systems: a perspective on redox systems biology. Free Radic. Biol. Med. 44, 921–937 73. Yoshida, K. et al. (2015) Thioredoxin selectivity for thiol-based redox regulation of target proteins in chloroplasts. J. Biol. Chem. 290, 4278–14288 74. Berndt, C. et al. (2015) The specificity of thioredoxins and glutaredoxins is determined by electrostatic and geometric complementarity. Chem. Sci. 6, 7049–7058 75. Palde, P.B. and Carroll, K.S. (2015) A universal entropy-driven mechanism for thioredoxin–target recognition. Proc. Natl. Acad. Sci. U. S. A. 112, 7960–7965 76. Yoshida, K. et al. (2014) Distinct redox behaviors of chloroplast thiol enzymes and their relationships with photosynthetic electron transport in Arabidopsis thaliana. Plant Cell Physiol. 55, 1415–1425 77. Rey, P. et al. (2013) Overexpression of plastidial thioredoxins f and m differentially alters photosynthetic activity and response to oxidative stress in tobacco plants. Front. Plant Sci. 4, 390 78. Nikkanen, L. and Rintamäki, E. (2014) Thioredoxin-dependent regulatory networks in chloroplasts under fluctuating light conditions. Philos. Trans. R. Soc. Lond. B 369, 20130224

Trends in Plant Science, Month Year, Vol. xx, No. yy

13

TRPLSC 1699 No. of Pages 14

79. Thormählen, I. et al. (2017) Thioredoxins play a crucial role in dynamic acclimation of photosynthesis in fluctuating light. Mol. Plant 10, 168–182

93. Grabsztunowicz, M. et al. (2017) Post-translational modifications in regulation of chloroplast function: recent advances. Front. Plant Sci. 8, 240

80. Yoshida, K. and Hisabori, T. (2017) Distinct electron transfer from ferredoxin-thioredoxin reductase to multiple thioredoxin isoforms in chloroplasts. Biochem. J. 474, 1347–1360

94. Venne, A.S. et al. (2014) The next level of complexity: crosstalk of posttranslational modifications. Proteomics 14, 513–524

81. Yoshida, K. and Hisabori, T. (2016) Two distinct redox cascades cooperatively regulate chloroplast functions and sustain plant viability. Proc. Natl. Acad. Sci. U. S. A. 113, 3967–3976

95. Nussinov, R. et al. (2012) Allosteric post-translational modification codes. Trends Biochem. Sci. 37, 447–455 96. Navrot, N. et al. (2011) Plant redox proteomics. J. Proteom. 74, 1450–1462

82. Nikkanen, L. et al. (2016) Crosstalk between chloroplast thioredoxin systems in regulation of photosynthesis. Plant Cell Environ. 39, 1691–1705

97. Roos, G. et al. (2015) How the disulfide conformation determines the disulfide/thiol redox potential. J. Biomol. Struct. Dyn. 33, 93–103

83. Geigenberger, P. et al. (2017) The unprecedented versatility of the plant thioredoxin system. Trends Plant Sci. 22, 249–262

98. Nagy, P. (2013) Kinetics and mechanisms of thiol-disulfide exchange covering direct substitution and thiol oxidation-mediated pathways. Antioxid. Redox Signal. 18, 1623–1641

84. Gütle, D.D. et al. (2017) Dithiol disulphide exchange in redox regulation of chloroplast enzymes in response to evolutionary and structural constraints. Plant Sci. 255, 1–11 85. König, N. and Scheibe, R. (2010) Redox-modification of chloroplast enzymes in Galdieria sulphuraria: trial-and-error in evolution or perfect adaptation to extreme conditions? In Red Algae in the Genomic Age (Seckbach, J. and Chapman, D.J., eds), pp. 391–407, Springer 86. Chiadmi, M. et al. (1999) Redox signalling in the chloroplast: structure of oxidized pea fructose-1,6-bisphosphate phosphatase. EMBO J. 18, 6809–6815 87. Johansson, K. et al. (1999) Structural basis of light activation of a chloroplast enzyme: the structure of sorghum NADP-malate dehydrogenase in its oxidized form. Biochemistry 38, 4319–4326 88. Carr, P.D. et al. (1999) Chloroplast NADP-malate dehydrogenase: structural basis of light-dependent regulation by thiol oxidation and reduction. Structure 7, 461–475 89. Maeda, K. et al. (2008) Crystal structures of barley thioredoxin h isoforms HvTrxh1 and HvTrxh2 reveal features involved in protein recognition and possibly in discriminating the isoform specificity. Protein Sci. 17, 1015–1024 90. Qin, J. et al. (2000) Molecular views of redox regulation: threedimensional structures of redox regulatory proteins and protein complexes. Antioxid. Redox Signal. 2, 827–840 91. Jung, Y.J. et al. (2013) Analysis of Arabidopsis thioredoxin-h isotypes identifies discrete domains that confer specific structural and functional properties. Biochem. J. 456, 13–24 92. Netto, L.E. et al. (2016) Conferring specificity in redox pathways by enzymatic thiol/disulfide exchange reactions. Free Radic. Res. 50, 206–245

14

Trends in Plant Science, Month Year, Vol. xx, No. yy

99. Poole, L.B. (2015) The basics of thiols and cysteines in redox biology and chemistry. Free Radic. Biol. Med. 80, 148–157 100. Jensen, K.S. et al. (2009) Kinetic and thermodynamic aspects of cellular thiol-disulfide redox regulation. Antioxid. Redox Signal. 11, 1047–1058 101. Zhu, L. et al. (2008) The torsin-family AAA+ protein OOC-5 contains a critical disulfide adjacent to Sensor-II that couples redox state to nucleotide binding. Mol. Biol. Cell 19, 3599–3612 102. Paulsen, C.E. and Carroll, K.S. (2010) Orchestrating redox signaling networks through regulatory cysteine switches. ACS Chem. Biol. 5, 47–62 103. Klomsiri, C. et al. (2011) Cysteine-based redox switches in enzymes. Antioxid. Redox Signal. 14, 1065–1077 104. Lee, S. et al. (2012) Thioredoxin and thioredoxin target proteins: from molecular mechanisms to functional significance. Antioxid. Redox Signal. 18, 1165–1207 105. Araki, K. et al. (2016) Redox sensitivities of global cellular cysteine residues under reductive and oxidative stress. J. Proteome. Res. 15, 2548–2559 106. Hancock, J. and Whiteman, M. (2018) Cellular redox environment and its influence on redox signaling molecules. Reactive Oxygen Species 5, 78–85 107. Bechtel, T.J. and Weerapana, E. (2017) From structure to redox: the diverse functional roles of disulfides and implications in disease. Proteomics 17, 1–21