ArticlePDF Available

Towards complete cofactor arrangement in the 3.0 A resolution structure of photosystem II

Authors:

Abstract and Figures

Oxygenic photosynthesis in plants, algae and cyanobacteria is initiated at photosystem II, a homodimeric multisubunit protein-cofactor complex embedded in the thylakoid membrane. Photosystem II captures sunlight and powers the unique photo-induced oxidation of water to atmospheric oxygen. Crystallographic investigations of cyanobacterial photosystem II have provided several medium-resolution structures (3.8 to 3.2 A) that explain the general arrangement of the protein matrix and cofactors, but do not give a full picture of the complex. Here we describe the most complete cyanobacterial photosystem II structure obtained so far, showing locations of and interactions between 20 protein subunits and 77 cofactors per monomer. Assignment of 11 beta-carotenes yields insights into electron and energy transfer and photo-protection mechanisms in the reaction centre and antenna subunits. The high number of 14 integrally bound lipids reflects the structural and functional importance of these molecules for flexibility within and assembly of photosystem II. A lipophilic pathway is proposed for the diffusion of secondary plastoquinone that transfers redox equivalents from photosystem II to the photosynthetic chain. The structure provides information about the Mn4Ca cluster, where oxidation of water takes place. Our study uncovers near-atomic details necessary to understand the processes that convert light to chemical energy.
Content may be subject to copyright.
© 2005 Nature Publishing Group
Towards complete cofactor arrangement in the
3.0 A
˚
resolution structure of photosystem II
Bernhard Loll
1
*†, Jan Kern
2
*, Wolfram Saenger
1
, Athina Zouni
2
& Jacek Biesiadka
1
Oxygenic photosynthesis in plants, algae and cyanobacteria is
initiated at photosystem II, a homodimeric multisubunit protein–
cofactor complex embedded in the thylakoid membrane
1
. Photo-
system II captures sunlight and powers the unique photo-induced
oxidation of water to atmospheric oxygen
1,2
. Crystallographic
investigations of cyanobacterial photosystem II have provided
several medium-resolution structures (3.8 to 3.2 A
˚
)
3–6
that explain
the general arrangement of the protein matrix and cofactors, but
do not give a full picture of the complex. Here we describe the most
complete cyanobacterial photosystem II structure obtained so far,
showing locations of and interactions between 20 protein subunits
and 77 cofactors per monomer. Assignment of 11
b
-carotenes
yields insights into electron and energy transfer and photo-
protection mechanisms in the reaction centre and antenna sub-
units. The high number of 14 integrally bound lipids reflects the
structural and functional importance of these molecules for
flexibility within and assembly of photosystem II. A lipophilic
pathway is proposed for the diffusion of secondary plastoquinone
that transfers redox equivalents from photosystem II to the
photosynthetic chain. The structure provides information
about the Mn
4
Ca cluster, where oxidation of water takes place.
Our study uncovers near-atomic details necessary to understand
the processes that convert light to chemical energy.
In photosystem II (PSII), excitation energy is transferred from the
antenna system to the reaction centre, where the primary electron
donor P680 formed by chlorophyll a (Chla) molecules is excited to
P680*, followed by release of an electron that travels along the
electron transfer chain by means of the pheophytin a Pheo
D1
to the
plastoquinone Q
A
, forming P680
þ
Q
A
2
. After two steps of reduction
and protonation of the secondary plastoquinone Q
B
, the plasto-
quinol PQH
2
leaves the Q
B
site and is replaced by new plastoquinone
PQ (ref. 1). Oxidation of water at the Mn
4
Ca cluster occurs in four
steps (termed S states) in the so-called Kok cycle
2
; at each step the
water-oxidizing complex is oxidized to a higher oxidation state, and
after the fourth step molecular O
2
is released. The electrons are
transferred from the Mn
4
Ca cluster through redox-active Tyr
Z
to
P680
þ
, which is reduced to P680 for another photosynthetic
cycle. Our structure of dimeric PSII from Thermosynechococcus
elongatus shows the location of 35 Chla,11
b
-carotene (Car), two
pheophytin (Pheo), two PQ, two haem, bicarbonate, 14 lipid and
three n-dodecyl-
b
-D-maltoside (
b
-DM) detergent molecules, the
Mn
4
Ca cluster, and one Fe
2þ
and one putative Ca
2þ
ion in each
monomer. The cofactor and protein arrangement is shown in Fig. 1
and Supplementary Fig. S1.
Various studies have indicated the influence of lipids on the
assembly and function of PSII
7,8
, but so far no structural information
about lipids bound to PSII is available. We found 14 lipids integrally
bound to PSII: four DGDG (digalactosyldiacylglycerol), six MGDG
(monogalactosyldiacylglycerol), three SQDG (sulphoquinovosyl-
diacylglycerol) and one PG (phosphatidyldiacylglycerol) molecule
(Fig. 1 and Supplementary Fig. S2). The composition of the lipids is
comparable to that of the thylakoid membrane
9
. Hydrophilic lipid
headgroups are close to the membrane surface, whereas hydrophobic
fatty acid chains are anchored between transmembrane
a
-helices
(TMHs) of different subunits. Negatively charged headgroups of
SQDG and PG are exclusively located at the cytoplasmic side
(Fig. 1b).
A belt of 11 lipids surrounds the reaction centre, separating it from
the antenna and smaller subunits (Fig. 2a). The remaining three
lipids and the detergent molecules are located at the monomer–
monomer interface. The unusually high lipid content in the PSII
complex provides a structural flexibility that might be required for
local mobility of subunits and promotes subunit–subunit recog-
nition. At high light intensities, Chla triplet states can be populated
and react with molecular oxygen to form singlet oxygen, leading to
oxidative damage of proteins. As subunit D1 is most prone to this
photodamage, it needs to be continuously replaced by newly syn-
thesized D1 (ref. 10). To facilitate this turnover of D1, a flexible
environment, such as that provided by the belt of lipids, might be
necessary. A similar ‘lubricant’ role may be fulfilled by lipid molecules
located at the dimerization interface, as observed in other multimeric
membrane-intrinsic complexes
11
.
The protein environment of Q
A
,Q
B
and non-haem-Fe
2þ
is
described in the Supplementary Discussion and Fig. S3. The
Q
B
-binding site (Fig. 2b) opens into a large cavity defined by
the TMHs of cytochrome b-559 (PsbE and PsbF), PsbJ, PsbK, TMH-d
and TMH-e of D1, and TMH-a of D2 (Fig. 2 and Supplementary
Fig. S4). The walls of the cavity are lipophilic as they are coated by
phytol chains of P
D2
, Chl
D2
, Pheo
D2
, Chla37, Chla44 and Chla46,
and by acyl chains of four lipids. In addition, Car11 and Car12 point
into the cavity. Fragmented electron density in the cavity is not
interpretable and could be due to two or three disordered lipophilic
molecules.
The cavity has two openings perpendicular to each other: the
larger one (,16 £ 16 A
˚
) opens towards the cytoplasmic side and the
smaller one (,10 £ 20 A
˚
), flanked by TMHs of cytochrome b-559
and PsbJ, faces the membrane interior (Fig. 2c). On the opposite side
of the membrane-facing opening, the Q
B
-binding pocket forms an
antechamber of the cavity that is filled by the quinone headgroup,
while its isoprenoid chain winds along the wall of the cavity. This
arrangement suggests that the cavity provides a flexible, lipophilic
environment for the diffusion of PQ/PQH
2
through the membrane-
facing opening, between the Q
B
-binding site in PSII and the plasto-
quinone pool of the thylakoid membrane (Fig. 2d). Similar lipophilic
LETTERS
1
Institut fu¨r Chemie und Biochemie/Kristallographie, Freie Universita
¨
t Berlin, Takustrasse 6, D-14195 Berlin, Germany.
2
Institut fu¨r Chemie/Max Volmer Laboratorium fu¨r
Biophysikalische Chemie, Technische Universita
¨
t Berlin, Straße des 17 Juni 135, D-10623 Berlin, Germany. †Present address: Max-Planck-Institut fu¨r Medizinische Forschung,
Abteilung fu¨r Biomolekulare Mechanismen, D-69120 Heidelberg, Germany.
*These authors contributed equally to this work.
Vol 438|15 December 2005|doi:10.1038/nature04224
1040
© 2005 Nature Publishing Group
pathways have been proposed for the photosynthetic cytochrome b
6
f
complex
12,13
and for the respiratory cytochrome bc
1
complex
14
. The
alternative exchange pathway for PQ through the larger cytoplasmic
opening is less favourable because hydrophobic PQ molecules would
have to pass through the cytosol before entering the membrane. It is
likely that PsbJ and cytochrome b-559, flanking the membrane-
facing opening, could regulate PQ diffusion. Indeed, PsbJ has been
shown to influence the electron transfer between Q
A
,Q
B
and the
Figure 2 | The plastoquinone diffusion pathway.
a, Lipid molecules located in PSII, viewed as in
Fig. 1. Lipid, detergent and quinone molecules are
shown in space-filling representation, and
a
-helices of protein subunits are shown as
cylinders. Subunits D1 and D2 and cytochrome
b-559 are coloured as in Fig. 1, the remaining
a
-helices are grey, and lipids are cyan. TMHs
forming the wall of the cavity are labelled.
b,Q
B
-binding pocket. The interacting residues of
D1 are labelled, arrows indicate or ientations of
TMH-d and TMH-e, and of the connecting
a
-helix
de. c, Electrochemical potential surface of PSII
(red, negative; blue, positive; white, neutral),
viewed along the membrane plane onto the cavity
opening that faces the membrane. The tail of Q
B
(violet) is nestled in the cavity. d, Schematic views
of Q
B
diffusion pathway in two orientations. Top,
same view as a. The positions of TMHs forming the
wall of the cavity (circles) and the membrane-
facing opening are indicated. Car molecules are
orange; lipid molecules are black. Bottom, PSII
embedded in the membrane. Positions and
approximate dimensions of the two openings of
the cavity, as well as the position of the Q
B
-binding
pocket and lipid molecules, are indicated.
Figure 1 | The PSII monomer from the cytoplasmic side. a, Overall view of
the PSII monomer. Transmembrane
a
-helices are represented as cylinders;
other protein elements are omitted for clarity. The main subunits are
highlighted as follows: reaction centre subunits D1 (yellow) and D2
(orange), antenna subunits CP43 (magenta) and CP47 (red), and the
a
- and
b
-chain of cytochrome b-559 (green and cyan, respectively). Low molecular
mass subunits are coloured grey. Unassigned TMHs are labelled X1–X3.
Cofactors are coloured green (Chla), yellow (Pheo), red (Car), blue (haem),
violet (quinone), black (lipids) and brown (detergent and unassigned alkyl
chains). The non-haem Fe
2þ
(blue) and putative Ca
2þ
(yellow) ions are
shown as spheres. Dashed line indicates the monomer–monomer interface.
b, Schematic representation of the view in a, with the same colour code.
Lipid and detergent molecules with headgroups pointing ‘downwards or
‘upwards are located at the lumenal or cytop lasmic side, respectively. Green
numbers in CP43 and CP47 indicate the positions of antenna Chla. The Q
B
diffusion cavity is indicated by a dotted line.
NATURE|Vol 438|15 December 2005 LETTERS
1041
© 2005 Nature Publishing Group
plastoquinone pool in tobacco
15
and, because the haem of cyto-
chrome b-559 is exposed to the cavity, an influence of the interior of
the cavity on the redox properties of cytochrome b-559 can also be
imagined.
The 11 carotenoid molecules were modelled as
b
-carotene in
all-trans configuration. Two carotenoids are bound to the D1 and
D2 subunits (Fig. 3), three are at CP43 and five at CP47 (Fig. 1).
Car12 is nestled between subunits PsbJ, PsbK, PsbZ and X1. The
carotenoids are distributed uniformly at the periphery of PSII except
for Car3, Car4, Car5 and Car6, which are clustered at the monomer–
monomer interface. This contrasts with the arrangement of seven
Car molecules in previous work
5
(Supplementary Discussion). In the
reaction centre, Car
D2
(Car11) is nearly parallel to the membrane
plane and located close to cytochrome b-559, as shown in previous
models
5,6
. Its counterpart Car
D1
(Car1) is oriented roughly perpen-
dicular to the membrane plane, the isoprenoid moiety being parallel
to TMH-a of D1 (Fig. 3a and Supplementary Fig. S5). The presence
of two differently oriented and therefore spectroscopically distinct
Car molecules has been shown for spinach PSII (ref. 16), where Car
D1
corresponds to Car
489
and Car
D2
to Car
507
.
Because Car
D1
does not bridge between Chlz
D1
and other cofactors
of the electron transfer chain (Fig. 3b), it is unlikely that Car
D1
participates in electron transfer reactions. Its position is rather
optimized for transfer and/or quenching of Chlz
D1
triplet states
and quenching of singlet oxygen that could be produced by
3
P680
(ref. 17). By contrast, the position of Car
D2
seems to be less efficient
for quenching of triplet states, but is in keeping with a role as a
‘molecular wire’ in putative secondary electron transfer when the
Mn
4
Ca cluster is not functional or is absent
17
. This idea is supported
by calculations of maximal electron transfer rates, showing that
secondary electron transfer reactions will predominantly occur on
the D2 side (Supplementary Table S4).
Most of the Car molecules located in the antenna subunits are in
van der Waals contact with Chla molecules, thereby allowing transfer
of excitation energy from Car to Chla, as well as protection of PSII
from destructive triplet states of Chla by means of rapid triplet–
triplet transfer. However, not all Chla molecules are located close to a
Car molecule. As coupled Chla molecules of the antenna might act as
excitation energy sinks, with a longer lifetime of the excited state and
hence higher probability of triplet formation, sufficient protection
from triplet states could be achieved if these groups are connected to
at least one Car molecule. Strongest coupled Chla groups in CP47
(Chla11, Chla12, Chla13 and Chla14, Chla17, Chla25, Chla26,
Chla27) have a Car molecule at van der Waals contact, whereas
corresponding groups in CP43 (Chla34, Chla46 and Chla45, Chla47)
are more distant from any Car, indicating lower triplet quenching
efficiency. A cluster of four Car molecules (Car3, Car4, Car5 and
Car6) is located at the monomer–monomer interface, close to a
coupled group of five Chla molecules in CP47 (Fig. 1 and Sup-
plementary Fig. S6). Three of the clustered Car molecules are at van
der Waals distance to each other, forming a coupled Car multimer,
which could broaden the absorption spectrum of the antenna system.
The shape of the electron density of the Mn
4
Ca cluster can be best
approximated by four Mn ions arranged as a ‘hook’ (Fig. 4a)
resembling a Y-shaped structure
3,6
. The Mn ions are numbered
Mn1 to Mn4, starting from the highest electron density at the bend
of the ‘hook’. The Mn–Mn distances could not be derived directly at
3.0 A
˚
resolution and thus were crystallographically refined with the
aid of restraints obtained from EXAFS studies (Supplementary
Methods). Mn1–Mn2 and Mn2–Mn3 are 2.7 A
˚
apart and probably
connected by di-
m
-oxo bridges, whereas Mn1–Mn3 and Mn3–Mn4
are at 3.3 A
˚
, suggesting mono-
m
-oxo bridges (Fig. 4b). This arrange-
ment is compatible with the ‘3 þ 1’ models suggested for the Mn
cations
18
. The position of Ca
2þ
between the Mn and Tyr
Z
is supported
by spherical anomalous difference electron density (Supplementary
Fig. S7). Ca
2þ
forming the vertex of a trigonal pyramid is equidistant
(,3.4 A
˚
) to Mn1, Mn2 and Mn3.
The interpretation of the Mn
4
Ca structure and coordination must
account for likely radiation damage of the cluster during X-ray data
collection. EXAFS studies of PSII crystals and solutions indicated
that Mn
3þ
and Mn
4þ
ions of the Mn
4
Ca cluster are rapidly reduced
by X-ray generated radicals to Mn
2þ
(ref. 19). This reduction is
probably associated with structural changes caused by the disruption
of
m
-oxo bridges and Mn–ligand interactions and may lead to
disorder, as indicated by uneven distribution of electron density in
the cluster (Supplementary Fig. S7).
The two short and two long Mn–Mn interactions and the three
equal Mn–Ca
2þ
interactions differ from the numbers of such
interactions derived from EXAFS studies
20
. This discrepancy could
be caused by the still low resolution of the diffraction data and by
possible radiation damage. In particular, the positions of Ca
2þ
and
Mn4 are less well defined in the electron density, leading to higher
coordinate errors for these atoms. The proposed coordination of Mn
and Ca
2þ
is shown in Fig. 4b. The residues of the second coordination
Figure 3 | Redox-active cofactors and electron transfer chain. a, View
along the membrane plane. The cofactors of the electron transfer chain
(P
D1
/P
D2
, Chl
D1
/Chl
D2
, Pheo
D1
/Pheo
D2
,Q
A
/Q
B
) are related by the pseudo-
C2 axis (arrow). Fe
2þ
(blue), Mn (red) and Ca
2þ
(yellow) ions are shown as
spheres. b, Schematic representation of the view in a. Selected distances
(in angstroms) are drawn between cofactor centres (black lines) and edges of
p
systems (red dotted lines).
LETTERS NATURE|Vol 438|15 December 2005
1042
© 2005 Nature Publishing Group
sphere (Fig. 4b, grey), even though they are not close enough to
coordinate the Mn
4
Ca cluster directly in the oxidation states present
in the crystals, might be suitable for interaction through water
molecules or may provide ligation in different redox states of the
Mn
4
Ca cluster.
Although the limited resolution does not allow a definite decision
about rotamers of ligating side chains, five carboxylates are in
positions that could act as bidentate ligands bridging different
cations. The combination of bidentate ligands and mono- or di-
m
-
oxo bridges could enhance the stability of the cation arrangement
and facilitate rearrangement of
m
-oxo bridges during the S-state
cycle. Mutational studies of ligating residues
21,22
found that, except
for Glu 333 and Asp 342 of D1, replacement of a ligating group with a
non-ligating group still leads to a (partially) functional cluster. Our
structure shows that in most cases a second ligating group is present,
which could hold the cations in place even when the exchanged
residue cannot coordinate them.
Glu 189 and Ala 344 in D1 are identified as possible ligands for
Ca
2þ
, with the former also ligating Mn1 and the latter Mn2, in
contrast to previous findings
5
. Taking into account the positional
error of Ca
2þ
and results from Fourier transform infrared (FTIR)
spectroscopy indicating that Ala 344 of D1 should ligate Mn
(ref. 23) but not Ca
2þ
(ref. 24), it is likely that Ca
2þ
is coordinated
only by Glu 189 of D1. Mn4, which is coordinated by Asp 170 and
Glu 333 of D1 and represents the first Mn ion bound to the high-
affinity binding site during assembly of the cluster
22,25
, seems to be
more distinct from the other three Mn ions. Our electron density
suggests that Mn4 is most prone to radiation damage and subsequent
disorder.
Our structure of the Mn
4
Ca cluster differs considerably from the
postulated cubane-like model
5
because the Mn–Mn distances in
the pyramid formed by three Mn and Ca
2þ
are not equal and the
pyramid is connected asymmetrically to Mn4 (Fig. 4c; for a detailed
comparison of both structures, see Supplementary Discussion and
Fig. S8). Even though the same coordinating residues have been
proposed
5
, none of them was modelled as bridging two metal ions,
leading to a lower saturation of the metal coordination spheres.
Consequently, the restraints for possible
m
-oxo-bridging are altered,
resulting in a different architecture of the whole oxygen-evolving
complex and providing new implications for the mechanism of water
oxidation.
Combining the structure with results from FTIR studies
23,26,27
and
assuming an oxidation state distribution of Mn(
III)
2
(IV)
2
for the S
1
state of the Kok cycle
28,29
, we can suggest localization of oxidation
states on individual metal ions of the cluster. Mn1 and Mn3 can
attain either oxidation state
III or oxidation state IV in the S
1
state.
Mn4, which is ligated by Asp 170, is not oxidized during S
0
–S
1
–S
2
–S
3
transitions
26
and is therefore probably present as Mn(IV), whereas
Mn2, which is ligated by the carboxy-terminal carboxylate of Ala 344,
changes from Mn(
III) to Mn(IV) in the S
1
–S
2
transition
23
and can be
considered as the site of first oxidation in the catalytic cycle of water
splitting.
METHODS
Dimeric PSII from T. elongatus was purified, characterized and crystallized as
described
30
. X-ray diffraction data were collected at the European Synchrotron
Radiation Facility (ESRF, Grenoble, France). The crystal structure of PSII was
determined to 3.0 A
˚
resolution and refined to R and R
free
factors of 0.24 and 0.29,
respectively (Supplementary Methods and Tables S1 and S2).
Received 1 July; accepted 13 September 2005.
1. Ke, B. Photosynthesis
Photobiochemistry and Photobiophysics (Kluwer Academic,
Dordrecht, 2001).
2. Kok, B., Forbush, B. & McGloin, M. Cooperation of charges in photosynthetic
evolution: I. A linear four step mechanism. Photochem. Photobiol. 11, 457–-475
(1970).
3. Zouni, A. et al. Crystal structure of photosystem II from Synechococcus
elongatus at 3.8 A
˚
resolution. Nature 409, 739–-743 (2001).
4. Kamiya, N. & Shen, J. R. Crystal structure of oxygen-evolving photosystem II
from Thermosynechococcus vulcanus at 3.7-A
˚
resolution. Proc. Natl Acad. Sci.
USA 100, 98–-103 (2003).
5. Ferreira, K. N., Iverson, T. M., Maghlaoui, K., Barber, J. & Iwata, S. Architecture
of the photosynthetic oxygen-evolving center. Science 303, 1831–-1838
(2004).
6. Biesiadka, J., Loll, B., Kern, J., Irrgang, K.-D. & Zouni, A. Crystal structure of
cyanobacterial photosystem II at 3.2 A
˚
resolution: a closer look at the Mn-
cluster. Phys. Chem. Chem. Phys. 6, 4733–-4736 (2004).
7. Gombos, Z. et al. Phosphatidylglycerol requirement for the function of electron
acceptor plastoquinone Q
B
in the photosystem II reaction center. Biochemistry
41, 3796–-3802 (2002).
8. Steffen, R., Kelly, A. A., Huyer, J., Do
¨
rmann, P. & Renger, G. Investigations on
the reaction pattern of photosystem II in leaves from Arabidopsis thaliana wild
type plants and mutants with genetically modified lipid content. Biochemistry
44, 3134–-3142 (2005).
9. Wada, H. & Murata, N. in Lipids in Photosynthesis: Structure, Function and
Genetics (ed. Murata, N.) 65–-81 (Kluwer Academic, Dordrecht, 1998).
10. Baena-Gonzalez, E. & Aro, E. M. Biogenesis, assembly and turnover of
photosystem II units. Phil. Trans. R. Soc. Lond. B 357, 1451–-1459 (2002).
11. Palsdottir, H. & Hunte, C. Lipids in membrane protein structures. Biochim.
Biophys. Acta 1666, 2–-18 (2004).
12. Stroebel, D., Choquet, Y., Popot, J. L. & Picot, D. An atypical haem in the
cytochrome b
6
f complex. Nature 426, 413–-418 (2003).
Figure 4 | Oxygen-evolving centre. a, Electron density of the Mn
4
Ca
cluster, viewed similarly to Fig. 3a. Surrounding amino acids of D1 (yellow)
and CP43 (magenta) are indicated; Mn is red and Ca
2þ
is orange. The
2F
o
2 F
c
map is contoured at the 1.2j level. b, Schematic view of the Mn
4
Ca
cluster. Distances between Mn (red) and Ca
2þ
(orange) are indicated by the
connecting lines (grey, 2.7 A
˚
; blue, 3.3 A
˚
, green, 3.4 A
˚
). Amino acids of the
first coordination sphere are black; those of the second sphere are grey;
distances are given in angstroms. c, Superposition of the current Mn
4
Ca
structure (coloured as in a) with that of ref. 5 (D1 in green, CP43 in orange,
Mn in violet and Ca
2þ
in yellow, respectively).
NATURE|Vol 438|15 December 2005 LETTERS
1043
© 2005 Nature Publishing Group
13. Kurisu, G., Zhang, H., Smith, J. L. & Cramer, W. A. Structure of the cytochrome
b
6
f complex of oxygenic photosynthesis: tuning the cavity. Science 302,
1009–-1014 (2003).
14. Lange, C., Nett, J. H., Trumpower, B. L. & Hunte, C. Specific roles of protein–-
phospholipid interactions in the yeast cytochrome bc
1
complex structure.
EMBO J. 20, 6591–-6600 (2001).
15. Ohad, I., Dal Bosco, C., Herrmann, R. G. & Meurer, J. Photosystem II proteins
PsbL and PsbJ regulate electron flow to the plastoquinone pool. Biochemistry
43, 2297–-2308 (2004).
16. Kwa, S. L. S., Newell, W. R., van Grondelle, R. & Dekker, J. P. The reaction
center of Photsystem II studied with polarized fluorescence spectroscopy.
Biochim. Biophys. Acta 1099, 193–-202 (1992).
17. Telfer, A. What is b-carotene doing in the photosystem II reaction centre? Phil.
Trans. R. Soc. Lond. B 357, 1431–-1440 (2002).
18. Peloquin, J. M. & Britt, R. D. EPR/ENDOR characterization of the physical and
electronic structure of the OEC Mn cluster. Biochim. Biophys. Acta 1503, 96–-111
(2001).
19. Yano, J. et al. X-ray damage to the Mn
4
Ca complex in single-crystals of
photosystem II: a case study for metallo-protein X-ray crystallography. Proc.
Natl Acad. Sci. USA 102, 12047–-12052 (2005).
20. Sauer, K., Yano, J. & Yachandra, V. K. X-ray spectroscopy of the Mn
4
Ca cluster
in the water-oxidation complex of photosystem II. Photosynth. Res. 85, 73–-86
(2005).
21. Diner, B. A. Amino acid residues involved in the coordination and assembly of
the manganese cluster of photosystem II. Proton-coupled electron transport of
the redox-active tyrosines and its relationship to water oxidation. Biochim.
Biophys. Acta 1503, 147–-163 (2001).
22. Debus, R. J. Amino acid residues that modulate the properties of tyrosine Y(Z)
and the manganese cluster in the water oxidizing complex of photosystem II.
Biochim. Biophys. Acta 1503, 164–-186 (2001).
23. Kimura, Y., Mizusawa, N., Yamanari, T., Ishii, A. & Ono, T. A. Structural
changes of D1 C-terminal
a
-carboxylate during S-state cycling in
photosynthetic oxygen evolution. J. Biol. Chem. 280, 2078–-2083 (2005).
24. Strickler, M. A., Walker, L. M., Hillier, W. & Debus, R. J. Evidence from
biosynthetically incorporated strontium and FTIR difference spectroscopy that
the C-terminus of the D1 polypeptide of photosystem II does not ligate
calcium. Biochemistry 44, 8571–-8577 (2005).
25. Campbell, K. D. et al. Dual-mode EPR detects the initial intermediate in
photoassembly of the photosystem II Mn cluster: the influence of amino acid
residue 170 of the D1 polypeptide on Mn coordination. J. Am. Chem. Soc. 122,
3754–-3761 (2000).
26. Debus, R. J., Strickler, M. A., Walker, L. M. & Hillier, W. No evidence from FTIR
difference spectroscopy that aspartate-170 of the D1 polypeptide ligates a
manganese ion that undergoes oxidation during the S
0
to S
1
,S
1
to S
2
,orS
2
to
S
3
transitions in photosystem II. Biochemistry 44, 1367–-1374 (2005).
27. Kimura, Y., Ishii, A., Yamanari, T. & Ono, T. A. Water-sensitive low-frequency
vibrations of reaction intermediates during S-state cycling in photosynthetic
water oxidation. Biochemistry 44, 7613–-7622 (2005).
28. Roelofs, T. A. et al. Oxidation states of the manganese cluster during the
flash-induced S-state cycle of the photosynthetic oxygen-evolving complex.
Proc. Natl Acad. Sci. USA 93, 3335–-3340 (1996).
29. Dau, H., Iuzzolino, L. & Dittmer, J. The tetra-manganese complex of
photosystem II during its redox cycle
X-ray absorption results and
mechanistic implications. Biochim. Biophys. Acta 1503, 24–-39 (2001).
30. Kern, J. et al. Purification, characterisation and crystallisation of photosystem II
from Thermosynechococcus elongatus cultivated in a new type of
photobioreactor. Biochim. Biophys. Acta 1706, 147–-157 (2005).
Supplementary Information is linked to the online version of the paper at
www.nature.com/nature.
Acknowledgements We thank D. DiFiore and C. Lu¨neberg for technical
assistance; K.-D. Irrgang, H. Lokstein, J. Messinger, F. Mu¨h, T. Renger,
E. Schlodder and J. Yano for discussions; and R. Clarke, G. Renger, K. Sauer and
V. Yachandra for critically reading the manuscript. Beam time and support at
the ESRF, SLS, BESSY and DESY are gratefully acknowledged. We thank
Deutsche Forschungsgemeinschaft and Fonds der Chemischen Industrie for
support.
Author Information Atomic coordinates have been deposited in the Protein
Data Bank under the accession number 2AXT. Reprints and permissions
information is available at npg.nature.com/reprintsandpermissions. The authors
declare no competing financial interests. Correspondence and requests for
materials should be addressed to W.S. (saenger@chemie.fu-berlin.de) and A.Z.
(zouni@phosis1.chem.tu-berlin.de).
LETTERS NATURE|Vol 438|15 December 2005
1044
... Ces protéines sont codées par des gènes psbC et psbB du génome chloroplastique, chacune possède six hélices α transmembranaires (Bricker, 1990 (Loll et al., 2005). CP47 étant localisé adjacent à la protéine D2 interagit plus directement et fortement avec le centre réactionnel du PSII que CP43 qui est adjacent à la protéine D1 (Bricker and Frankel, 2002 (Amunts et al., 2007). ...
... Les sites de l'OEC situés sur l'interface lumenale de la membrane sont entourés par quatre unités protéiques (D1, D2, CP43 et CP47), mais presque tous les ligands de cluster Mn4Ca viennent de la protéine D1 avec un seul ligand provenant de CP43. Selon des études par diffraction des rayons X publiées plus ou moins récemment(Fereirra et al., 2004;Loll et al., 2005;Guskov et al., 2009), trois ions Mn (Mn1, Mn2 et Mn3) avec un ion Ca placé dans le sommet sont groupé dans une pyramide trigonale, avec le quatrième atome de Mn (Mn4) asymétriquement connecté à l'un des coins de la base. Ca 2+ est un élément structurel essentiel dans le cluster métal et est aussi probablement d'être un site de liaison (connexion) de molécules d'H2O substrats qui subit l'oxydation. ...
Thesis
Full-text available
Effet du stress thermique sur le transport cyclique des électrons autour du PSI dans des mutants d'Arabidopsis thaliana déficients dans certaines voies cycliques (crr2-2 et pgr5) et dans d'autres mutants défectueux en digalactosyle-diacylglycerole, DGDG (dgd1-2 et dgd1-3)
... Both triplet chlorophylls and singlet oxygen then extract hydrogen from unsaturated lipids, initiating lipid peroxidation [56][57][58]. The rapid and prolonged production of ROS overwhelms the quenching capacity of photoprotective components near PS II [59,60], such as carotenoids, xanthophylls, tocopherol, and flavonoids. This imbalance leads to severe oxidative damage to proteins, lipids, and pigments, ultimately resulting in cell membrane destruction and plant death [61]. ...
Article
Full-text available
Utilizing Unmanned Aerial Vehicle (UAV) multispectral technology offers a non-destructive and efficient approach to monitoring plant health and stress responses by analyzing reflectance data across various wavelengths. This study integrates UAV-based multispectral imagery with ground-measured sample data to evaluate the impact of atrazine (ATR) on chlorophyll a, chlorophyll b, carotenoids, and anthocyanins in Red Cos lettuce. The results indicate a significant increase (p < 0.05) in ATR concentration in lettuce with soil application, leading to notable reductions in pigment concentrations. Heatmap analysis reveals that EVI shows the strongest negative correlations with pigment classes (coefficients ranging between -0.75 to -0.85), while NDVI, GNDVI, and BNDVI exhibit the strongest positive correlations with pigments (coefficients >0.75). These findings highlight the potential of this innovative technique in predicting pigment concentrations and emphasize its importance in monitoring pesticide effects for sustainable agriculture.
... PSII absorbs light and then breaks down water, providing electrons to PSI while generating H + , oxygen, and ATP, which is used in the Calvin cycle for carbon fixation. 36 Figure 6B shows the changes in the oxygen release rate of Chlorella PY-ZU1 under different culture temperatures. The pattern of photosynthetic oxygen release was similar to that of biomass dry weight accumulation, with the 35°C and 37.5°C temperature conditions exceeding the others during the same period, and expressing a decreasing trend in the oxygen release rate over time. ...
Article
Full-text available
To explore the effects of waste heat (50–170°C) from steel plant flue gas on the column photobioreactor algal liquid temperature for microalgal production, a flue gas‐microalgal liquid heat transfer model was developed that simulated the microalgal growth environment for flue‐gas carbon dioxide (CO2) fixation. The simulation results showed that the influence of high‐temperature flue gas weakened with the increasing microalgal liquid temperature due to enhanced evaporation and heat dissipation. Increasing the flue gas temperature and aeration rate resulted in a higher microalgal liquid temperature up to a maximum increase of 4.16°C at an ambient temperature of 25°C, an aeration rate of 2 L/min, and a flue gas temperature of 170°C. In an experiment on the effect of incubation temperature on the growth rate of microalgae, at an optimal temperature of 35°C, the Chlorella sp. PY‐ZU1 growth rate exhibited a remarkable increase of 104.7% compared to that at 42.5°C. Therefore, modulating the flue gas conditions can significantly increase the microalgal growth rate for CO2 fixation, making it a promising approach to increase biomass production for efficient carbon utilization.
... These lipids can play both structural and functional roles, as highlighted in the literature [81]. For instance, a study on the cyanobacterium Thermosynechococcus elongatus revealed that several lipids are found in the photosystem II (PSII) and appear to have a role in the organization and functionality of this photosynthetic structure, with MGDG, DGDG, SQDG, and PG present around the D1/D2 reaction center of PSII [82,83]. Among these lipids, SQDG, an anionic lipid, has been shown to be particularly important for the growth and photosynthesis of T. elongatus. ...
Article
Full-text available
Abstract The cultivation of cyanobacteria by exploiting available in situ resources represents a possible way to supply food and oxygen to astronauts during long-term crewed missions on Mars. Here, we evaluated the possibility of cultivating the extremophile cyanobacterium Chroococcidiopsis thermalis CCALA 050 under operating conditions that should occur within a dome hosting a recently patented process to produce nutrients and oxygen on Mars. The medium adopted to cultivate this cyanobacterium, named Martian medium, was obtained using a mixture of regolith leachate and astronauts’ urine simulants that would be available in situ resources whose exploitation could reduce the mission payload. The results demonstrated that C. thermalis can grow in such a medium. For producing high biomass, the best medium consisted of specific percentages (40%vol) of Martian medium and a standard medium (60%vol). Biomass produced in such a medium exhibits excellent antioxidant properties and contains significant amounts of pigments. Lipidomic analysis demonstrated that biomass contains strategic lipid classes able to help the astronauts facing the oxidative stress and inflammatory phenomena taking place on Mars. These characteristics suggest that this strain could serve as a valuable nutritional resource for astronauts.
... These hexagonal structures can serve to stabilize membrane proteins under stressful conditions [53]. DGDGs were localized mainly in the reaction centre of PSII [66,68]. Our results seem to demonstrate that the natural conversion of MGDG to DGDG improves under stress conditions. ...
Article
Full-text available
Abstract In the present work, the possibility to grow the strain Synechococcus nidulans CCALA 188 on Mars using a medium mimicking a one obtainable using in situ available resources, i.e. the so-called Martian medium, under an atmosphere obtainable by pressurization of Mars CO2, is investigated. The goal is to obtain a biomass with high-value products to sustain a crewed mission to Mars. The results show that the replacement of 40% vol of Z-medium with the same volume of Martian medium does not afect the cultivation and leads to a slight improvement of biomass productivity. Under an atmosphere consisting of pure CO2 the growth rate was reduced but the strain managed to adapt by modifying its metabolism. Total proteins and carbohydrates were signifcantly reduced under Mars-like conditions, while lipids increased when using CO2. A balanced diet rich in antioxidants is crucial for the wealth of astronauts, and in our case, radical scavenging capacities range from 15 to 20 mmolTEAC/kg were observed. Under CO2, a reduction in antioxidant power is observed likely due to a decrease in photosynthetic activity. The lipidome consisted of sulfoquinovosyldiacylglycerol, monogalactosyldiacylglycerol, digalactosyldiacylglycerol, phosphatidylcholine, phosphatidylglycerol, and triacylglycerol. A signifcant increase in the latter ones was observed under Mars simulated atmosphere.
... MGDG is involved in a number of critical photosynthetic activities, such as photoprotection, controlling pigment composition and nonphotochemical quenching, and producing a proton motive force across the thylakoid membrane in high light (Aronsson et al. 2008). Because it influences the activities of both PSI and PSII, DGDG is essential for preserving the best possible efficiency of photosynthetic electron transport (Hartel et al. 1997;Liu et al. 2004;Loll et al. 2005). In the plant galactolipid skeleton, the most prevalent fatty acids in approximate proportions of 80%, 16%, 16%, and 70% are 18:3/16:3 as 34:6-MGDG, 18:3/18:3 as 36:6-MGDG, 18:3/16:0 as 34:3-MGDG, and 18:3/18:3-DGDG (Chen et al. 2010). ...
Article
Full-text available
The membranes of plants are where temperature sensing begins and where freezing injury typically occurs. Barley’s adaptation to and survival after freezing stress is aided by remodelling of its membrane lipid composition. The modifications of individual lipid molecular species in different stress-treated plant species and cultivars can indicate the functions of genes regulating lipid metabolism or signaling. In this study, we employed a membrane lipidomic approach to investigate the response of barley of two cold-tolerant and two cold-sensitive cultivars to freezing temperatures during the barley trefoil stage. A total of 56 predominant lipid compounds changed significantly under freezing stress were identified. Phosphatidic acid (PA), lysophosphatidic acid (LPA) and monogalactosyldiacylglycerol (MGDG) in freezing-tolerant varieties were significantly upregulated under freezing stress, while there was a decrease in freezing-sensitive cultivars. Freezing-tolerant varieties experienced greater changes in lipid composition compared to freezing-sensitive cultivars, which had proportionally smaller changes. In addition, when exposed to short-term cold stress, varieties A and B had lower levels of monoglyceride lipase (MGLL) than varieties C and D. However, under long-term cold stress, the opposite was observed. Additionally, the freezing-tolerant variety A showed a notable increase in the expression of diacylglycerol acyltransferase 1 (DGAT1) after being exposed to 4 °C. Furthermore, SENSITIVE TO FREEZING 2 (SFR2) reached its highest level in all four varieties after being exposed to cold treatment for 48 h. This study indicates that freezing injury in barley leaves is correlated with extensive changes in lipid metabolism and that freezing-tolerant varieties can alleviate freezing injury by membrane lipid remodelling. The study’s outcomes may improve our understanding of barley’s freezing adaptation mechanisms and contribute to breeding for better tolerance.
Chapter
More than ever before in human history, climate change, which is largely caused by humans, will pose serious existential challenges for the first time. In particular, the future energy supply of mankind will play a decisive role. Fossil fuels can no longer be a future option unless the risk of further excessive use is ignored and accepted. There are sufficient alternatives available to replace fossil fuels. Renewable energies, including biomass, can replace almost 100 % of fossil fuels in an intelligent mix of wind, solar, tidal, and geothermal energy. In addition, efficient and optimised interaction is possible when combined with artificial intelligence. This chapter focuses on the use of biological processes and their contribution to future energy management and highlights synergies and trade-offs. The aim is to make it clear that with the help of these technologies, mankind can certainly keep the complex Earth system from becoming completely out of balance.
Chapter
Various extreme habitats such as thermal springs, deserts, hypersaline environment, polar regions, and acidic pools are present in the Earth, and these habitats harbor extremophilic oxygenic photosynthetic organisms. Photosynthetic organisms play key roles in sustaining the ecosystems of extremophilic habitats by acting as primary producers. Phototrophs thriving in these environments may face multiple extreme factors at the same time, e.g., in acidic thermal springs, both acidic conditions and high temperature are available. Phototrophs inhabiting extreme habitats exhibit a wide range of diversity, i.e., from prokaryotic archaea, bacteria, and cyanobacteria to eukaryotic algae and plants, which have been discussed in the present chapter. To sustain the photosynthetic processes at extreme habitats, these phototrophs possess some special adaptation strategies including morphological alterations, modulations in light absorption machinery, changes in carbon fixation enzymes, and structural and compositional modifications in chloroplasts and cyanobacteria. While the morphological changes, especially in the leaf structure and canopy architecture, govern the survival of higher plants in extreme conditions, adaptations in the composition and structure of phycobiliproteins form the basis of cyanobacterial existence in such conditions. Other notable adaptive responses of cyanobacteria thriving in the extreme environments include substitution of some amino acids in the proteins of photosystem II (PSII) complex and plastocyanin. Sequence analysis of D1 proteins of PSII from different groups of extremophilic phototrophs reflected the occurrence of many isoforms of this protein. Nevertheless, such analysis of plastocyanin synthesized by extremophilic phototrophs has not been performed yet. In this chapter, phylogenetic analysis of plastocyanin showed that extremophilic plants and eukaryotic algae share a similar type of plastocyanin, while the plastocyanin of extremophilic cyanobacteria made separate clade. Besides, the chapter also encompasses elaborate discussion on the pellucid role of compatible solutes and alkanes for unabating photosynthesis in extremophilic phototrophs. Although many research articles, related to photosynthesis behavior of organisms in extreme conditions, have been found in the literature, comprehensive compilation of these studies is still needed to further elucidate the existing lacunae. So, this chapter inclusively has diversity of extremophilic photosynthetic organisms and their adaptation features which make their photosynthesis sustainable in extreme conditions.
Article
Full-text available
Photosystem II (PSII), the principal enzyme of oxygenic photosynthesis, contains two integral light harvesting proteins (CP43 and CP47) that bind chlorophylls and carotenoids. The two intrinsic antennae play crucial roles in excitation energy transfer and photoprotection. CP43 interacts most closely with the reaction center of PSII, specifically with the branch of the reaction center (D1) that is responsible for primary charge separation and electron transfer. Deciphering the function of CP43 requires detailed atomic-level insights into the properties of the embedded pigments. To advance this goal, we employ a range of multiscale computational approaches to determine the site energies and excitonic profile of CP43 chlorophylls, using large all-atom models of a membrane-bound PSII monomer. In addition to time-dependent density functional theory (TD-DFT) used in the context of a quantum-mechanics/molecular-mechanics setup (QM/MM), we present a thorough analysis using the perturbed matrix method (PMM), which enables us to utilize information from long-timescale molecular dynamics simulations of native PSII-complexed CP43. The excited state energetics and excitonic couplings have both similarities and differences compared with previous experimental fits and theoretical calculations. Both static TD-DFT and dynamic PMM results indicate a layered distribution of site energies and reveal specific groups of chlorophylls that have shared contributions to low-energy excitations. Importantly, the contribution to the lowest energy exciton does not arise from the same chlorophylls at each system configuration, but rather changes as a function of conformational dynamics. An unexpected finding is the identification of a low-energy charge-transfer excited state within CP43 that involves a lumenal (C2) and the central (C10) chlorophyll of the complex. The results provide a refined basis for structure-based interpretation of spectroscopic observations and for further deciphering excitation energy transfer in oxygenic photosynthesis.
Article
Full-text available
The influence of poly(ethylene glycol) (PEG) polymers H–(O–CH2–CH2)p–OH with different average molecular sizes p\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$p$$\end{document} on the micelle formation of n-alkyl-β-D-maltoside detergents with the number of carbon atoms in the alkyl chain ranging from 10\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$10$$\end{document} to 12\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$12$$\end{document} is investigated with the aim to learn more about the detergent behavior under conditions suitable for the crystallization of the photosynthetic pigment–protein complex photosystem II. PEG is shown to increase the critical micelle concentration (CMC) of all three detergents in the crystallization buffer in a way that the free energy of micelle formation increases linearly with the concentration of oxyethylene units (O–CH2–CH2) irrespective of the actual molecular weight of the polymer. The CMC shift is modeled by assuming for simplicity that it is dominated by the interaction between PEG and detergent monomers and is interpreted in terms of an increase of the transfer free energy of a methylene group of the alkyl chain by 0.2 kJ mol⁻¹ per 1 mol L⁻¹ increase of the concentration of oxyethylene units at 298 K. Implications of this effect for the solubilization and crystallization of protein–detergent complexes as well as detergent extraction from crystals are discussed.
Book
Photosynthesis: Photobiochemistry and Photobiophysics is the first single-authored book in the Advances in Photosynthesis Series. It provides an overview of the light reactions and electron transfers in both oxygenic and anoxygenic photosynthesis. The scope of the book is characterized by the time frame in which the light reactions and the subsequent electron transfers take place, namely between <=10⁻¹² and >=10⁻³ second. The book is divided into five parts: An Overview; Bacterial Photosynthesis; Photosystem II & Oxygen Evolution; Photosystem I; and Proton Transport and Photophosphorylation. In discussing the structure and function of various protein complexes, we begin with an introductory chapter, followed by chapters on light-harvesting complexes, the primary electron donors and the primary electron acceptors, and finally the secondary electron donors. The discussion on electron acceptors is presented in the order of their discovery to convey a sense of history, in parallel with the advancement in instrumentation of increasing time resolution. The book includes a large number of stereo pictures showing the three-dimensional structure of various photosynthetic proteins, which can be easily viewed with unaided eyes. This book is designed to be used as a textbook in a graduate or upper-division undergraduate course in photosynthesis, photobiology, plant physiology, biochemistry, and biophysics; it is equally suitable as a resource book for students, teachers, and researchers in the areas of molecular and cellular biology, integrative biology, microbiology, and plant biology.
Article
In the crystal structure of photosystem II (PSII) from the cyanobacterium Thermosynechococcus elongatus at 3.2 Å resolution, several loop regions of the principal protein subunits are now defined that were not interpretable previously at 3.8 Å resolution. The head groups and side chains of the organic cofactors of the electron transfer chain and of antenna chlorophyll a (Chl a) have been modeled, coordinating and hydrogen bonding amino acids identified and the nature of the binding pockets derived. The orientations of these cofactors resemble those of the reaction center from anoxygenic purple bacteria, but differences in hydrogen bonding and protein environment modulate their properties and provide the unique high redox potential (1.17 V) of the primary donor. Coordinating amino acids of manganese cluster, redox-active TyrZ and non-haem Fe2+ have been determined, and an all-trans β-carotene connects cytochrome b-559, ChlZ and primary electron donor (coordinates are available under PDB-code 1W5C).
Article
— Using isolated chloroplasts and techniques as described by Joliot and Joliot[6] we studied the evolution of O2 in weak light and light flashes to analyze the interactions between light induced O2 precursors and their decay in darkness. The following observations and conclusions are reported: 1. Light flashes always produce the same number of oxidizing equivalents either as precursor or as O2. 2. The number of unstable precursor equivalents present during steady state photosynthesis is ∼ 1.2 per photochemical trapping center. 3. The cooperation of the four photochemically formed oxidizing equivalents occurs essentially in the individual reaction centers and the final O2 evolution step is a one quantum process. 4. The data are compatible with a linear four step mechanism in which a trapping center, or an associated catalyst, (S) successively accumulates four + charges. The S4+ state produces O2 and returns to the ground state S0. 5. Besides S0 also the first oxidized state S+ is stable in the dark, the two higher states, S2+ and S3+ are not. 6. The relaxation times of some of the photooxidation steps were estimated. The fastest reaction, presumably S*1←S2, has a (first) half time ≤ 200 μsec. The S*2 state and probably also the S*0 state are processed somewhat more slowly (˜ 300–400 μsec).
Article
We report the first parallel polarization EPR signal from the Mn(III) ion formed by photooxidation of Mn(II) bound at the high affinity Mn-binding site of photosystem II (PSII). This species corresponds to the first photoactivation intermediate formed on the pathway to assembly of the water-splitting Mn cluster. The parallel mode EPR spectrum of the photooxidation product of 1.2/1 stoichiometry Mn(II)/Mn-depleted wild-type Synechocystis sp. PCC 6803 PSII particles consists of six well-resolved transitions split by a relatively small 55Mn hyperfine coupling (44 G). This spectral signature is absent in photooxidized Mn apoPSII complexes prepared from D1-Asp170Glu and D1-Asp170His mutants, providing direct spectral evidence for a role for this specific D1-Asp170 residue in the initial photoactivation chemistry. Temperature-dependence measurements and spectral simulations performed on the Mn(III) parallel mode EPR signal of the wild-type sample give an axial zero-field splitting value of D ≈ −2.5 cm-1 and a rhombic zero-field splitting value of |E| ≈ 0.269 cm-1. The negative D value for this d4 ion is indicative of either a 5B1g symmetry ground state of an octahedral Mn(III) geometry or a 5B1 symmetry ground state of a five-coordinate square-pyramidal Mn(III) geometry. The parallel mode Mn(III) EPR spectrum obtained from the wild-type photooxidized Mn apoPSII complex is contrasted with that obtained from the five-coordinate Mn(III) form of native Mn superoxide dismutase, which has a trigonal-bipyramidal geometry and a 5A1 symmetry ground state giving rise to a positive D value and a much larger 55Mn hyperfine coupling of 100 G. The D1-Asp170His mutant displays a parallel mode EPR spectrum similar to that observed in a Mn(III) model complex. The D1-Asp170Glu mutant shows no parallel mode spectrum, but in perpendicular mode it shows a broad feature near g = 5 which has spectral characteristics of an S = 3/2 Mn(IV) ion. This suggests that this mutant provides a binding site with a less positive Mn(III)/Mn(IV) reduction potential.
Article
Low-temperature steady-state emission properties have been analyzed of Photosystem II reaction center (RC) complexes isolated from spinach CP47-RC complexes after a short Triton X-100 treatment and stabilization in n-dodecyl β,d-maltoside. Excitation spectra of the fluorescence anisotropy were detected at the maximum of the single fluorescence band at 683.5 nm and at the vibrational subband of the same emission at 742 nm. The Q1 transitions of the red-most absorbing pigment(s) showed positive anisotropy with a value of about 0.22. The value is lower than that of the theoretical maximum (0.4) and is explained by a combination of (1) vibrational depolarization effects and (2) by assuming that the red-most absorbing pigments arise from the low-exciton component of P680, that the exciton coupling breaks upon excitation, and that the angle between the monomer Q1 transitions of P680 is 48 ± 10°. The Q1 transitions of pheophytin showed negative anisotropy. This result, combined with the results obtained with linear dichroism spectroscopy, suggests that the spatial organization of the Q1 transitions of pheophytin matches the organization of the bacteriopheophytin residues in the bacterial reaction center. The spatial organization of the y-polarized transitions of pheophytin could be similar in both systems, although these transitions could also be tilted somewhat more towards the membrane plane in PS II. The data furthermore indicate that the accessory chlorophylls in PS II and in bacterial reaction centers have different average orientations, and suggest that at least some of the accessory chlorophylls in PS II have a pheophytin-like orientation.
Article
The catalytic site for photosynthetic water oxidation is embedded in a protein matrix consisting of nearly 30 different polypeptides. Residues from several of these polypeptides modulate the properties of the tetrameric Mn cluster and the redox-active tyrosine residue, Y(Z), that are located at the catalytic site. However, most or all of the residues that interact directly with Y(Z) and the Mn cluster appear to be contributed by the D1 polypeptide. This review summarizes our knowledge of the environments of Y(Z) and the Mn cluster as obtained from the introduction of site-directed, deletion, and other mutations into the photosystem II polypeptides of the cyanobacterium Synechocystis sp. PCC 6803 and the green alga Chlamydomonas reinhardtii.
Article
The combination of site-directed mutagenesis, isotopic labeling, new magnetic resonance techniques and optical spectroscopic methods have provided new insights into cofactor coordination and into the mechanism of electron transport and proton-coupled electron transport in photosystem II. Site-directed mutations in the D1 polypeptide of this photosystem have implicated a number of histidine and carboxylate residues in the coordination and assembly of the manganese cluster, responsible for photosynthetic water oxidation. Many of these are located in the carboxy-terminal region of this polypeptide close to the processing site involved in its maturation. This maturation is a required precondition for cluster assembly. Recent proposals for the mechanism of water oxidation have directly implicated redox-active tyrosine Y(Z) in this mechanism and have emphasized the importance of the coupling of proton and electron transfer in the reduction of Y(Z)(radical) by the Mn cluster. The interaction of both homologous redox-active tyrosines Y(Z) and Y(D) with their respective homologous proton acceptors is discussed in an effort to better understand the significance of such coupling.
Article
Electron paramagnetic resonance (EPR) spectroscopy has often played a crucial role in characterizing the various cofactors and processes of photosynthesis, and photosystem II and its oxygen evolving chemistry is no exception. Until recently, the application of EPR spectroscopy to the characterization of the oxygen evolving complex (OEC) has been limited to the S2-state of the Kok cycle. However, in the past few years, continuous wave-EPR signals have been obtained for both the S0- and S1-state as well as for the S2 (radical)(Z)-state of a number of inhibited systems. Furthermore, the pulsed EPR technique of electron spin echo electron nuclear double resonance spectroscopy has been used to directly probe the 55Mn nuclei of the manganese cluster. In this review, we discuss how the EPR data obtained from each of these states of the OEC Kok cycle are being used to provide insight into the physical and electronic structure of the manganese cluster and its interaction with the key tyrosine, Y(Z).