Sie sind auf Seite 1von 12

Skip to Main content

Elsevier logo

Photosystem I

Related terms:

ChloroplastChlorophyllThylakoidPhotosystemPhotosystem IIFerredoxinPhotonElectron Transport

View all Topics

Organellar and Metabolic Processes

Kevin E. Redding, in The Chlamydomonas Sourcebook, 2009

A. Function of photosystem I

Photosystem I (PS I) is a chlorophyll (Chl)–protein complex that functions as a light-driven


plastocyanin:ferredoxin oxidoreductase. Electron transfer from plastocyanin (Em ≈ +370 mV) to
ferredoxin (Em ≈ −430 mV) would normally be very endergonic (ΔG ≈ +87 kJ/mol), but is rendered
favorable by coupling to absorption of a photon of visible light. For example, the energy of a red photon
(λ=700 nm) is ∼170 kJ/mol, which is more than sufficient to drive the otherwise-unfavorable reaction.
Thus, PS I can be thought of as a light-driven electron pump, transferring electrons from plastocyanin (or
cytochrome c6) on the lumenal side to ferredoxin on the stromal side, both across the thylakoid
membrane and over an energy barrier. PS I can function as part of the linear or cyclic electron transport
pathways.

Introduction to Chlamydomonas and its Laboratory Use

In The Chlamydomonas Sourcebook, 2009

F. Photosystem I

Photosystem I receives electrons from plastocyanin or cytochrome c6 on the lumenal side of the
thylakoid membrane and uses light energy to transfer them across the membrane to ferredoxin on the
stromal side. It can also function in a cyclic electron transport pathway. The core structure is a
heterodimer of the PsaA and PsaB proteins, which are encoded by chloroplast genes. As in Photosystem
II, light is harvested by antenna complexes, and the primary light reaction is a charge separation
beginning stabilized by transfer of an electron to a quinone, but in Photosystem I the terminal electron
acceptor is an FeS cluster, which permits reduction of ferredoxin. The peripheral subunits PsaC, PsaD,
and PsaE form the docking site for ferredoxin. The primary electron donor, P700, is ultimately reduced
by plastocyanin or cytochrome c6. PsaF and PsaJ are required for docking these proteins. Details of the
molecular structure are provided in Volume 2, Chapter 15.

The mechanism of energy transfer and trapping in Photosystem I

H.M. Vaswani, ... G.R. Fleming, in Femtochemistry and Femtobiology, 2004

3. THE ENERGY LANDSCAPE OF PHOTOSYSTEM I

Photosystem I represents a classic example of a system where high quality structural information,
though essential, is inadequate to understand the dynamical behavior of a biological system. The
absorption spectrum of PSI is very broad compared to the spectrum of Chl in solution. The dense
packing revealed in Figure 1 can be expected to produce a broad range of spectral shifts through
intermolecular interactions. In addition, the Chls are bound in 96 non-equivalent sites each with their
own unique set of polar, hydrogen bonding, and macrocycle distortion interactions. Our solution to this
problem was to calculate the electronic excitation energies of all 96 Chls, including their neighboring
protein residues and use these results to construct a 96 × 96 Hamiltonian for the complex. This is then
used to calculate the low temperature spectrum of the entire complex, and then coupled with a spectral
density taken from experiment [46–48], to calculate the ambient temperature spectrum[49–51]. The
calculated excitation energies are summarized in Figure 2 in which excitation energies are plotted as a
function of distance from Chl EC-A1, one of the two Chls that constitute the primary electron donor,
P700. A very broad distribution of energies is evident, but there is clearly no evidence for a downhill
(funnel) energy landscape in the bulk antenna.

Fig. 2. Calculated Qy excitation energies of Chls in PSI and its relation to the distance of the Chl from Chl
EC-A1, one of the two Chls in the special pair, P700.

PHOTOSYNTHETIC GENERATORS OF PROTONMOTIVE FORCE

In Bioenergetics 2, 1992

6.4.3 Photosystem I
Review Golbeck and Bryant, 1991

The understanding of photosystem I (PSI) is less securely based than that of PSII. PSI contains chlorophyll
as the pigment absorbing at 700 nm (P700), and it very probably but not definitely (at the time of
writing) functions as a dimer, as is known for the bacterial special pair and as is postulated for PSII.
There is suggestive evidence that there may be significant analogies between the organizations of the
two photosystems (Nitschke and Rutherford, 1991). The Em,7 of unexcited P700 is about + 450mV.
Following excitation of P700 the electron arrives within 10 ps on a chlorophyll a molecule known as A0.
After a further 100 ps the electron passes to A1 which is generally believed to be phylloquinone, vitamin
K1 (Fig 6.8). The electron ultimately reaches FeS centres located within the complex (Fig. 6.8).
Presumably one of these serves as the electron donor to reduce the iron–sulphur protein ferredoxin in
the aqueous phase. Bound ferredoxin has an Em,7 of − 530 mV, and so is extremely electronegative. The
carotenoid band shift (Section 6.3), which was first detected in chloroplasts, indicates that the electron
is transferred across the membrane in less than 20 ns. At the other end of the complex, there appears to
be a direct electron transfer from plastocyanin to P700+.

Figure 6.8. A plausible model for Photosystem I.

The model is heavily based on analogy to the bacterial reaction centre. There is no definite proof (but
some evidence is in favour) for a special pair of chlorophylls as P700 nor for the adjacent monomeric
chlorophylls ((Chi)?). A0 and A1 are widely considered to be chlorophyll and phylloquinone; Fx, FA and
FB are iron–sulphur centres. Some of the polypeptides are shown schematically with their molecular
weights and gene names. Based on Nitschke and Rutherford (1991) and Golbeck and Bryant (1991).

Whereas photosystem II (PSII) appears to be closely related to the reaction centre of R. sphaeroides and
R. viridis, there are grounds for supposing that the reaction centre in green sulphur bacteria and
heliobacteria may be closely related to photosystem I (PSI) (Nitschke and Rutherford, 1991). An
important difference between the reaction centres in at least some green bacteria and purple bacteria is
that photochemistry in the former produces a species with a much more reducing, i.e. negative, Em
value, to the extent that in some species NAD(P)+ can be reduced directly. Thus in these green bacteria
electrons from donors such as sulphide can be added at quinone and exit from reaction centre to a
component, probably a ferredoxin, that can reduce NAD(P)+ without any requirement for Δp-driven
reversed electron transport (see Gregory, 1989).
Mitochondria, Chloroplasts, Peroxisomes

In Cell Biology (Third Edition), 2017

Oxygen-Producing Synthesis of NADPH and ATP by Dual Photosystems

Chloroplasts and cyanobacteria combine photosystem II and photosystem I in the same membrane to
form a system capable of accepting low-energy electrons from the oxidation of water and producing
both a proton gradient to drive ATP synthesis and reducing equivalents in the form of NADPH (Fig.
19.8E–F). Both photosystems are more elaborate in dual systems than in single systems. Although plant
photosystem II, with more than 25 protein subunits, is much more complicated than is the homologous
reaction center of purple bacteria, the arrangement of transmembrane helices and chlorophyll cofactors
in the core of the plant reaction center is similar to the simple reaction center of purple bacteria.

Photosynthesis involves a tortuous electron transfer pathway powered at two waystations by


absorption of photons. This process begins when the special pair chlorophylls of photosystem II are
excited by direct absorption of light or by resonance energy transfer from surrounding light-harvesting
complexes (Fig. 19.8E–F). Electrons come from splitting two waters into molecular oxygen and four
protons. Excited-state electrons tunnel through the redox cofactors and combine with protons from the
stroma (or cytoplasm in bacteria) to reduce quinone QB to QH2, a high-energy electron donor. QH2
diffuses to complex b6-f, the chloroplast equivalent of the mitochondrial bc1 complex. Passage of
electrons through complex b6-f releases protons from QH2 into the thylakoid lumen (or bacterial
periplasm), contributing to the proton gradient across the membrane.

Complex b6-f donates electrons from QH2 to photosystem I. Direct absorption of 680-nm light or
resonance energy transfer from surrounding light-harvesting complexes boosts special pair chlorophyll
electrons to a very high-energy, excited state (Fig. 19.8F). Excited-state electrons pass through
chlorophyll and iron-sulfur centers of photosystem I to the iron-sulfur center of the redox protein,
ferredoxin, on the cytoplasmic/stromal surface of the membrane. The enzyme nicotinamide adenine
dinucleotide phosphate (NADP) reductase combines electrons from ferredoxin with a proton to form
NADPH, the final product of this complex electron transfer pathway powered at two waystations by
absorption of photons. Uptake of stromal protons during NADPH formation contributes to the
transmembrane proton gradient for the synthesis of ATP. Antiporters in the inner membrane exchange
ATP for ADP, as in mitochondria.

The Regulatory Evaluation of the Skin Effects of Pesticides


Michael O’Malley, in Hayes' Handbook of Pesticide Toxicology (Third Edition), 2010

28.2.5.1 Bipyridyls

Bipyridyls (diquat and paraquat) disrupt photosystem I in photosynthesis (Paraquat Information Center,
2009) but have multiple sites of action in animal as well as plant cells.

Diquat

For 2007, California agricultural use data showed 3220 applications, for a total of 70,047 pounds used on
grains, nursery crops, potatoes, and rights of way and for landscape maintenance.

Physical Properties

Formula, C12H12N2.2Br; MW, 344.05; MP, 337°C; log P, −4.60; VP, < 1 × 10−7 mm Hg; solubility in H2O,
708,000 mg/l at 20°C; other solubilities: slightly soluble in alcohols and hydroxylic solvents; practically
insoluble in nonpolar organic solvents

Irritation Data

A liquid mixture of 2.3% diquat with oxyfluorfen, dicamba, and fluazifop-p-butyl caused severe corrosion
in the Draize assay. Three products (2.3–37.3% AI) caused moderate irritation and four (0.23–8.35%)
caused minimal irritation.

Sensitization Data

Five liquid products (2.3–37.2% AI) did not cause sensitization in the Buehler test.

California Illness Data

There were 22 cases associated with diquat in the handler database, principally involving direct
accidental contact with diquat. One of the sample cases shown in Table 28.6 involved failure to
decontaminate shoes after they became soaked with diquat and subsequent prolonged contact (case
1986-1498) similar to irritation of lower extremities caused by fumigants.
Paraquat

Paraquat is a contact herbicide and dessicant used to control weeds on a variety of grain, vegetable, and
fruit crops.

Physical Properties

Formula, C12H14N2; MW, 186; BP, at 760 mm Hg, decomposes at 175–180°C; log P, −4.22; VP,
approximately 0 mm Hg at 20°C; solubility in H2O, soluble in water; other solubilities: practically
insoluble in organic solvents

Irritation Data

Two liquid products (22.3 and 36.1% AI) caused moderate irritation in the Draize test. A 43.8% liquid
caused minimal irritation.

Skin injury associated with application of paraquat and with its misuse has been reported from many
areas of the world (Angelo et al., 1986; Botella et al., 1985; Cooper et al., 1994; Gamier et al., 1994;
George, 1989; Horiuchi and Ando, 1980; Howard, 1979; Li, 1986; Peachey, 1981; Sugaya, 1976; Swan,
1969; Vilaplana et al., 1993; Villa et al., 1995). Skin injury in most cases has not been associated with
systemic effects of paraquat but has occasionally been described.

Sensitization Data

Two liquid products (22.3 and 37.1% AI) caused no sensitization in the Buehler test.

California Illness Data

There were 35 cases associated with paraquat in the California handler database. The sample cases
described in Table 28.6 involved a mild reaction following direct contact with a dilute paraquat spray
(1983-480).

INHIBITOR AND PLASTOQUINONE BINDING TO PHOTOSYSTEM II

Walter Oettmeier, Achim Trebst, in The Oxygen Evolving System of Photosynthesis, 1983
I INTRODUCTION

The reducing power of photosystem II in photosynthetic electron transport is conferred via the
cytochrome b6/f-complex to photosystem I. Electron transfer between photosystem II and the
cytochrome b6/f-complex is mediated by plastoquinone. Amongst the other electron carriers of the
photosynthetic electron transport chain plastoquinone is unique in that respect that it not only shuttles
electrons but simultaneously protons across the thylakoid membrane during its catalytic cycle. Evidence
is accumulating that the proton/electron stoichiometry exceeds unity thus indicating a ‘Q or b-cycle’ in
the cytochrome b6/f-complex.

The elucidation of plastoquinone function has been greatly facilitated by investigations with specific
inhibitors. Some of these inhibitors play a dual role not only in displaying information on plastoquinone
function but are equally important as herbicides in crop protection.

Photosynthesis is an electron transport process

Hans-Walter Heldt, Birgit Piechulla, in Plant Biochemistry (Fourth Edition), 2011

The light energy driving the cyclic electron transport of PS I is only utilized for the synthesis of ATP

Besides the noncyclic electron transport discussed so far, cyclic electron transfer can also take place in
which the electrons from the excited photosystem I are transferred back to the ground state of PS I,
probably via the cyt-b6/f complex (Fig. 3.34). The energy thus released is used only for the synthesis of
ATP, and NADPH is not formed. This electron transport is termed cyclic photophosphorylation. In intact
leaves, and even in isolated intact chloroplasts, it is quite difficult to differentiate experimentally
between cyclic and non-cyclic photophosphorylation. It has been a matter of debate as to whether and
to what extent cyclic photophosphorylation occurs in a leaf under normal physiological conditions.
Recent evaluations of the proton stoichiometry of photophosphorylation (see section 4.4) suggest that
the yield of ATP in noncyclic electron transport is not sufficient for the requirements of CO2 assimilation,
and therefore cyclic photophosphorylation seems to be required to synthesize the lacking ATP.
Moreover, cyclic photophosphorylation must operate at very high rates in the bundle sheath
chloroplasts of certain C4 plants (section 8.4). These cells have a high demand for ATP and they contain
high PS I activity but very little PS II. Presumably, the cyclic electron flow is governed by the redox state
of the acceptor of the photosystem in such a way that by increasing the reduction of the NADP system,
and consequently of ferredoxin, the diversion of the electrons in the cycle is enhanced. The function of
cyclic electron transport is probably to adjust the rates of ATP and NADPH formation according to the
plant's demand.
Figure 3.34. Cyclic electron transport between photosystem I and the cyt-b6/f complex. The path of the
electrons from the excited PS I to the cyt-b6/f complex is still unclear.

Despite intensive investigations, the pathway of electron flow from PS I to the cyt-b6/f complex in cyclic
electron transport remains unresolved. It has been proposed that cyclic electron transport is structurally
separated from the linear electron transport chain in a super complex. Most experiments on cyclic
electron transport have been carried out with isolated thylakoid membranes that catalyze only cyclic
electron transport when redox mediators, such as ferredoxin or flavin adenine mononucleotide (FMN,
Fig. 5.16), have been added. Cyclic electron transport is inhibited by the antibiotic antimycin A. It is not
clear at which site this inhibitor functions. Antimycin A does not inhibit noncyclic electron transport.

Surprisingly, proteins of the NADP dehydrogenase complex of the mitochondrial respiratory chain
(section 5.5) have also been identified in the thylakoid membrane of chloroplasts. The function of these
proteins in chloroplasts is still not known. The proteins of this complex occur very frequently in
chloroplasts from bundle sheath cells of C4 plants, which have little PS II but a particularly high cyclic
photophosphorylation activity (section 8.4). These observations raise the possibility that in cyclic
electron transport the flow of electrons from NADPH or ferredoxin to plastoquinone proceeds via a
complex similar to the mitochondrial NADH dehydrogenase complex. As will be shown in section 5.5, the
mitochondrial NADH dehydrogenase complex transfers electrons from NADH to ubiquinone. Results
indicate that an additional pathway for a cyclic electron transport exists in which electrons are directly
transferred via a plastoquinone reductase from ferredoxin to plastoquinone.

Primary Nitrogen Metabolism

Peter J. Lea, in Plant Biochemistry, 1997

(b) Heterocyst metabolism

The heterocyst lacks photosystem II activity which ensures that no oxygen is liberated in the location of
the nitrogenase enzyme. The presence of photosystem I ensures that sufficient ATP can be generated by
cyclic photophosphorylation. The heterocysts also lack Rubisco and the Calvin cycle enzymes so that CO2
fixation is limited to the vegetative cells. Thus carbohydrate has to be transported into the heterocysts,
probably in the form of maltose. Glucose is oxidized via the pentose phosphate pathway to yield NADPH
in the reaction catalyzed by glucose-6-phosphate dehydrogenase (see Fig. 7.6).
Figure 7.6. Diagrammatic representation of carbon and nitrogen exchange between the heterocyst and
vegetative cells. Carbon compounds are transported from the vegetative cells into the nitrogen fixing
heterocyst in order to supply the necessary NADPH via the oxidative pentose pathway. The ammonia
produced by nitrogen fixation is assimilated by glutamine synthetase and the majority of glutamine
transported to the vegetative cell where it is converted into glutamate by glutamate synthase (GOGAT).
The heterocyst and vegetative cell thus demonstrate an early evolved form of cellular specialization
(Smith &amp; Gallon, 1993).

The ammonia formed by nitrogenase is immediately assimilated by glutamine synthetase, and the
majority is exported to the vegetative cells in the form of glutamine. Glutamate which is formed by
glutamate synthase is returned to the heterocyst to assimilate another molecule of ammonia (Hager et
al., 1983). Full details of the glutamate synthase cycle are given in section 7.4. The system described is
thus a rather elegant division of labor, in which CG*2 assimilation and nitrogen fixation are spatially
separated but at the same time intimately linked.

In the unicellular cyanobacterium Gloeothece, nitrogen fixation takes place at night and photosynthetic
CO2 assimilation and O2 evolution occur during the day. In Gloeothece, nitrogenase synthesis
commences two hours before entering the dark period. After the onset of darkness, both synthesis and
activity of nitrogenase are supported by ATP, reductant and carbon skeletons formed from the
metabolism of stored glucan reserves, accumulated during the light period. The newly fixed ammonia is
rapidly assimilated into amino acids. A fuller description of nitrogen fixation in the cyanobacterium is
given by Gallon & Chaplin (1988) and Gallon (1992).

Redox Control, Redox Signaling, and Redox Homeostasis in Plant Cells

Karl-Josef Dietz, in International Review of Cytology, 2003

2 Excitation Energy Dissipation and Distribution

Nonphotochemical energy quenching prevents overreduction of the photosystem II reaction centers.


Two major regulatory mechanisms regulate the diverting of light energy between photosystem II and
photosystem I on the one hand and between photochemistry, chlorophyll a fluorescence, and heat
dissipation on the other hand (Kruse, 2001). Proton accumulation in the thylakoid lumen induces a pH-
dependent type of nonphotochemical energy quenching (Horton et al., 1994). Violaxanthin deepoxidase
is activated at low lumen pH and converts violaxanthin to zeaxanthin. Binding of zeaxanthin to light-
harvesting complex (LHC) proteins increases nonradiative energy dissipation. Thereby, the excitation
pressure to the photosystems is relieved (Müller et al., 2001). Because ascorbate is the reductive
cosubstrate in the deepoxidation reaction, there is also a redox element involved in the reaction, but
apparently not as a redox regulatory mechanism (Fig. 5).

Fig. 5. Examples of redox regulation and redox signaling in the context of photosynthesis. The scheme
presents the photosynthetic electron transport chain as central element. The yellow lines symbolize
linear and cyclic electron flow. Red circles represent redox intermediates involved in redox regulation
and signaling. Processes regulated or affected by redox are (1) energy dissipation by the violaxanthin
cycle, (2) energy distribution among both photosystems by state transition, (3) Calvin cycle activity, (4)
chloroplast transcription and (5) translation, (6) export of pyridine nucleotides by activation of the
malate valve, and (7) nuclear gene expression. APx, ascorbate peroxidase; Asc, ascorbic acid; cPABP,
homologue to poly(A)-binding proteins; cPDI, chloroplast protein disulfide isomerase; Cyt, cytochrome;
Cyp, cyclophilin; Fd, ferredoxin; GSH, glutathione; LHC, light-harvesting complex; MDH, malate
dehydrogenase; PC, plastocyanin; PEP, plastome-encoded RNA polymerase; PQ, plastoquinone; Prx,
peroxiredoxin; PTK, plastid transcription protein kinase. Psbs and Pgr5 are two recently identified gene
products that affect PS II regulation and cyclic electron flow.

The intersystem electron transport from photosystem II to the cytochrome fb6 complex is mediated by
the mobile electron carrier plastohydroquinone (PQ). Following transfer of electrons to the cytochrome
b6f complex and release of H+ to the lumen, plastoquinone diffuses back to photosystem II and is
charged with electrons again. Thus, the PQ redox state reflects the relative activity of the photosystems
and the availability of terminal electron acceptors, with a highly reduced state corresponding to an
unbalanced high PS II activity or low electron demand. The relative activity of the photosystems can be
adjusted by optimizing the excitation energy. In this context, the redox state of PQ serves as signal for
state transition, namely from excitation state I with the LHC II being attached to photosystem II to state
II with LHC attachment to photosystem I. Biochemically, state transition is mediated by LHC threonine
phosphorylation. Titration of the LHC protein kinase resulted in a redox midpoint potential close to that
of PQ (Horton et al., 1981). As a consequence of phosphorylation, the trimeric LHC II monomerizes and
the monomers diffuse to the stroma thylakoids (Allen, 1992) where they attach to the PsaH subunit of
PS I (Lunde et al., 2000). In addition to regulation of LHC kinase by the redox state of plastoquinone,
Rintamäki and co-workers (2000) showed that LHC phosphorylation involves a thiol-based redox signal.
Thioredoxin reduces and inactivates the LHC II kinase. The mechanism allows the system to integrate
redox information from the acceptor site of PS I and the intersystem electron transport. State transition
will occur only when the electron pressure of both redox-sensing systems differs, indicating a
misbalance between excitation pressure of PS II and PS I. Snyders and Kohorn (1999) identified a family
of thylakoid-associated kinases (TAKs) that accept LHC as substrate for phosphorylation. A high
reduction state of the PQ pool correlates with phosphorylation of a set of other PS II reaction center
core polypeptides including PsbA (D1 protein), PsbB (CP43), PsbD, and PsbH (Kruse, 2001). At present
the redox-related phosphorylation of these PS II proteins may be part of the PS II repair cycle. The
reaction center protein D1 is subjected to a high rate of turnover, particularly under high light
conditions. The repair cycle involves successive disassembly of the reaction center ⧸light harvesting
complex, dimer-to-monomer conversion, translocation of the reaction center ⧸CP47 subcomplex to the
stroma thylakoids, and replacement of the D1 protein. Dephosphorylation of the D1 protein precedes
D1 degradation (Rintamäki et al., 1996). Thus redox-dependent phosphorylation ⧸dephosphorylation
could be regulating the kinetics of the PS II repair cycle.

Correlative evidence has been obtained that the plastoquinone pool is a dominating redox signal
regulating protein phosphorylation. The occupation of the plastoquinone-docking site Q0 of the
cytochrome b6f complex (Hamel et al., 2000) and the reduction state of the ferredoxin ⧸thioredoxin
system (Rintamäki et al., 2000) also affected the phosphorylation of thylakoid membrane proteins. Two
conclusions have to be drawn:

1.

1.

Apparently, interacting redox-dependent signaling pathways define the adaptive response of


photochemistry to changing conditions of excitation pressure and demand for reducing power.

2.

The specific nature of the redox signals and the signaling pathways is not understood and most
conclusions are based on correlative evidence at present.

Elsevier logo

About ScienceDirect

Remote access

Shopping cart

Advertise
Contact and support

Terms and conditions

Privacy policy

We use cookies to help provide and enhance our service and tailor content and ads. By continuing you
agree to the use of cookies.

Copyright © 2020 Elsevier B.V. or its licensors or contributors. ScienceDirect ® is a registered trademark
of Elsevier B.V.

Das könnte Ihnen auch gefallen