ArticlePDF Available

Rapid formation of the stable tyrosyl radical in photosystem II

Authors:

Abstract and Figures

Two symmetrically positioned redox active tyrosine residues are present in the photosystem II (PSII) reaction center. One of them, TyrZ, is oxidized in the ns-micros time scale by P680+ and reduced rapidly (micros to ms) by electrons from the Mn complex. The other one, TyrD, is stable in its oxidized form and seems to play no direct role in enzyme function. Here, we have studied electron donation from these tyrosines to the chlorophyll cation (P680+) in Mn-depleted PSII from plants and cyanobacteria. In particular, a mutant lacking TyrZ was used to investigate electron donation from TyrD. By using EPR and time-resolved absorption spectroscopy, we show that reduced TyrD is capable of donating an electron to P680+ with t1/2 approximately equal to 190 ns at pH 8.5 in approximately half of the centers. This rate is approximately 10(5) times faster than was previously thought and similar to the TyrZ donation rate in Mn-depleted wild-type PSII (pH 8.5). Some earlier arguments put forward to rationalize the supposedly slow electron donation from TyrD (compared with that from TyrZ) can be reassessed. At pH 6.5, TyrZ (t1/2 = 2-10 micros) donates much faster to P680+ than does TyrD (t1/2 > 150 micros). These different rates may reflect the different fates of the proton released from the respective tyrosines upon oxidation. The rapid rate of electron donation from TyrD requires at least partial localization of P680+ on the chlorophyll (PD2) that is located on the D2 side of the reaction center.
Content may be subject to copyright.
Rapid formation of the stable tyrosyl radical in
photosystem II
Peter Faller*, Richard J. Debus
†‡
, Klaus Brettel*, Miwa Sugiura
§
, A. William Rutherford*
, and Alain Boussac*
*Section de Bioe´ nerge´ tique, De´ partement de Biologie Cellulaire et Mole´ culaire (DBCM), Commissariat a` l’Energie Atomique (CEA) Saclay, Centre National
de la Recherche Scientifique (CNRS) Unite´ de Recherche Associe´ e (URA) 2096, 91191 Gif-sur-Yvette Cedex, France;
Department of Biochemistry,
University of California, Riverside, CA 92521; and
§
Department of Applied Biological Chemistry, Osaka Prefecture University,
1-1 Gakuen-cho, Sakai, Osaka 599-8531, Japan
Edited by Christopher T. Walsh, Harvard Medical School, Boston, MA, and approved October 12, 2001 (received for review July 24, 2001)
Two symmetrically positioned redox active tyrosine residues are
present in the photosystem II (PSII) reaction center. One of them,
TyrZ, is oxidized in the ns–
s time scale by P680
and reduced
rapidly (
s to ms) by electrons from the Mn complex. The other
one, TyrD, is stable in its oxidized form and seems to play no direct
role in enzyme function. Here, we have studied electron donation
from these tyrosines to the chlorophyll cation (P680
) in Mn-
depleted PSII from plants and cyanobacteria. In particular, a mutant
lacking TyrZ was used to investigate electron donation from TyrD.
By using EPR and time-resolved absorption spectroscopy, we show
that reduced TyrD is capable of donating an electron to P680
with
t
1/2
190 ns at pH 8.5 in approximately half of the centers. This rate
is 10
5
times faster than was previously thought and similar to the
TyrZ donation rate in Mn-depleted wild-type PSII (pH 8.5). Some
earlier arguments put forward to rationalize the supposedly slow
electron donation from TyrD (compared with that from TyrZ) can
be reassessed. At pH 6.5, TyrZ (t
1/2
2–10
s) donates much faster
to P680
than does TyrD (t
1/2
> 150
s). These different rates may
reflect the different fates of the proton released from the respec-
tive tyrosines upon oxidation. The rapid rate of electron donation
from TyrD requires at least partial localization of P680
on the
chlorophyll (P
D2
) that is located on the D2 side of the reaction
center.
T
yrosyl radicals play key roles in the mechanisms of a wide
range of enzymes. Understanding these roles and how pro-
teins are able to control these reactive species to carry out quite
specific chemical reactions has been an important aim of re-
searchers in this area (for review, see ref. 1). One of the earliest
demonstrations of redox active tyrosines was in photosystem II
(PSII), the water oxidizing enzyme, in which a tyrosyl radical,
designated tyrosine Z
(TyrZ
), is thought to play a key role in
the active site, abstracting electrons and possibly protons from
the substrate water that is bound to the highly oxidized Mn
cluster (2–5).
PSII contains a second tyrosyl radical, TyrD
, which is stable
during enzyme function and which is located in a 2-fold rota-
tionally symmetrical position to TyrZ on a subunit (D2) adjacent
to that (D1) in which water oxidation takes place. TyrZ and TyrD
are equally situated relative to the central chlorophyll (Chl) pair
(P
D1
and P
D2
) with the center-to-center distances for TyrZ–P
D1
and TyrD–P
D2
being 12.4 Å (6). It is this pair of Chls that bears
the photogenerated cation (P680
) that is considered to be the
oxidant for the tyrosines (refs. 6–12; for recent reviews see refs.
2 and 3). The existing enzyme may have evolved from an ancestor
in which the core was a homodimer with the two tyrosines having
identical redox functions (13).
Despite homologous positions in subunits D1 and D2, TyrZ
and TyrD exhibit extremely different kinetics, different redox
potentials, and play completely different functional roles in the
enzyme. Thus, they constitute an ideal system for understanding
how the protein environment controls the reactivity of these
species.
Several studies have focused on the differences in the kinetics
of formation and decay of the two radicals. Most of these studies
were conducted on Mn-depleted PSII because TyrZ
is much
longer lived in this material, decaying with t
1/2
values of 20600
ms, in contrast with oxygen-evolving PSII in which the t
1/2
values
are 0.03–1 ms (14–15). Compared with TyrZ
,TyrD
is very
stable; it decays with t
1/2
in the minutes–hours time range (refs.
16–18; for reviews, see refs. 2 and 3).
TyrD
is thought to be immobilized in a site with a well defined
and ordered H-bond interaction between its phenol oxygen and
a proton from D2 His-189 (in cyanobacteria; D2 His-190 in
higher plants) (19). Moreover, the site is well shielded from the
lumen (20–23) and is thought to be relatively hydrophobic. The
TyrZ
site is quite different. In the absence of Mn, TyrZ
is
readily accessible from the lumen (22–24) and is thought to be
in a more hydrophilic environment. In addition, TyrZ
exhibits
a much more disordered H-bond, and the ring seems to be more
mobile (25, 26). The hydrogen bond to TyrZ
seems to involve
D1 His-190 directly or indirectly (i.e., by means of one or more
water molecules) (for discussion, see refs. 2 and 3).
In the reduced state, TyrD and TyrZ are likely to be proton-
ated and to form a hydrogen bond to D2 His-189 and D1 His-190,
respectively (27, 28). For TyrZ, indications were obtained for a
minor population in which the hydrogen bonds to either water,
a hydroxylated amino acid side chain, or to the protonated
imidazole side chain of His (i.e., HisH
2
instead of HisH) (27).
Because TyrZ and TyrD are protonated in the reduced state
and deprotonated in the oxidized state (Tyr
), the oxidation
involves both electron and proton transfer. In PSII lacking the
Mn-cluster, the t
1/2
(190 ns–30
s) of the electron transfer from
TyrZ to P680
was reported to depend on the biological origin
of PSII, the presence of Ca
2
, and the pH (23, 24, 29–31). The
rate increased with the pH from tens of
s at acidic pH to
sor
sub-
s at basic pH, exhibiting pK
a
values of 8.3 (23), 7.0 (24), and
7.5 (31). This pK
a
was ascribed to the ionization of either TyrZ
or D1-His-190 (23, 24, 31). In the latter case, the suggestion is
that at low pH, the His is protonated and has to be deprotonated
before it is able to accept a proton from TyrZ. In contrast, at high
pH (8–9), the His is thought to be deprotonated already and,
therefore, the electron transfer can occur more rapidly (sub-
s
or
s) (31, 32).
It is widely accepted that electron donation from TyrD to
P680
is relatively slow compared with TyrZ, although its redox
potential is much lower (TyrD, 0.75 V; TyrZ, 0.95 V–1.1 V).
This notion is largely based on the work of Buser et al. (33), who
measured the yield of oxidation of all of the electron donors to
P680
at pH 5.0–7.5 in Mn-depleted PSII. A global fit gave a t
1/2
This paper was submitted directly (Track II) to the PNAS office.
Abbreviations: Mes, 2-(N-morpholino)ethanesulfonic acid; Chl, chlorophyll; P680, pho-
tooxidizable chlorophyll; P
D1
,P
D2
, two central chlorophylls bound to polypeptide D1 and
D2, respectively; PSII, photosystem II.
To whom reprint requests should be addressed. E-mail: Richard.Debus@ucr.edu or
rutherford@dsvidf.cea.fr.
The publication costs of this article were defrayed in part by page charge payment. This
article must therefore be hereby marked advertisement in accordance with 18 U.S.C.
§1734 solely to indicate this fact.
14368–14373
PNAS
December 4, 2001
vol. 98
no. 25 www.pnas.orgcgidoi10.1073pnas.251382598
1020 ms for the electron transfer from TyrD to P680
at pH
values between 5.0 and 7.5. Thus, the oxidation yield appeared
to be practically independent of pH over this range (33).
Intriguingly, indications for TyrD electron donation to P680
with a rate similar to TyrZ exist in the older literature, but these
have been largely overlooked by the field (34, 35). The present
work was performed to readdress the electron donation rate
from TyrD to P680
by using kinetic absorption and EPR
spectroscopies and a mutant lacking TyrZ.
Materials and Methods
Construction of D1-Y161F Mutants and Isolation of PSII Particles from
Synechocystis sp.
The D1-Y161F mutation was constructed in the
psbA-2 gene of the cyanobacterium Synechocystis sp. strain PCC
6803 (36). The plasmid bearing this mutation was transformed
into a host strain of Synechocystis that lacks all three psbA genes
(37) and contains a hexahistidine-tag (His-tag) fused to the C
terminus of CP47 (38). Single colonies were selected for their
ability to grow on solid media containing 5
gml kanamycin
monosulfate (37). The control wild-type strain was constructed
in identical fashion as the D1-Y161F mutant except that the
transforming plasmid carried no site-directed mutation. The
designation wild-type differentiates this strain from the native
wild-type strain that contains all three psbA genes and is
sensitive to antibiotics. Cells were propagated and PSII particles
were isolated as described (38). The PSII particles were isolated
in a buffer containing 50 mM 2-(N-morpholino)ethanesulfonic
acid (Mes)NaOH (pH 6.0), 20 mM CaCl
2
, 5 mM MgCl
2
, 0.03%
(wtvol) N-dodecyl
-D-maltoside, 25% (volvol) glycerol, and
concentrated to 1 mg of Chl per ml (38). The concentration of
Mes was decreased to 5 mM by 10-fold dilution and was followed
by reconcentration. The final solution contained 5 mM Mes
NaOH (pH 6.0), 20 mM CaCl
2
, 5 mM MgCl
2
, 0.03% (wtvol)
N-dodecyl
-D-maltoside, and 25% (volvol) glycerol. Mn-
depleted wild-type PSII particles were treated with NH
2
OH and
EDTA, as described (32). After concentration to 1mgofChl
per ml, the concentration of Mes in the Mn-depleted PSII
particles was decreased to 5 mM by dilution and reconcentration,
as described above. The mutant and wild-type PSII cores were
stored at 77 K at a concentration of about 1.5 mg Chl per ml.
After thawing, 20 mM Tris HCl (pH 8.5) was added to the
PSII. The PSII particles then were dark-adapted for2hatroom
temperature with 8 mM K
3
Fe
3
(CN)
6
8mMK
4
Fe
2
(CN)
6
.
These conditions result in an almost complete reduction of
TyrD. For time-resolved absorption measurements, the samples
were diluted to 0.15 mg Chl per ml in 15 mM NaCl0.3 M sucrose
and 50 mM of one of the following buffers: CAPSONaOH (pH
9.5), TrisHCl (pH 8.5 and pH 8.0), HepesNaOH (pH 7.5), or
MesNaOH (pH 6.5). For EPR, 40
l of concentrated PSII was
adjusted to pH 8.5 (pH 6.5) by adding 12.5
l of 0.4 M TrisHCl
at pH 8.5 (or MesNaOH pH 6.5) and then put into a small
volume (50
l) quartz flat cell.
EPR Spectroscopy. X-band EPR spectra and kinetics were re-
corded with an ESP 300 spectrometer (Bruker, Karlsruhe,
Germany) in a flat cell [small size for Synechocystis sp., normal
size (400
l) for spinach]. Excitation flashes were provided
with a Spectra-Physics GCR-23010 Nd-YAG laser (532 nm, 550
mJ, 8 ns).
Time-Resolved Absorption at Room Temperature. Excitation flashes
(300 ps, 532 nm) were provided with an Nd-YAG laser (Quantel,
Les Ulis, France) with an energy of about 1.5 mJcm
2
. The
measuring light was a continuous laser diode emitting at 820 nm
(SDL-5411-G1, Spectra Diode Labs, San Jose, CA). A cut-off
filter RG780 was placed before the cuvette and an 820-nm
interference filter (10-nm bandwidth) after the cuvette (path
length 10 mm, 2 mm broad). The measuring-light intensity and
the flash-induced changes were monitored behind the sample
with a photodiode (FND 100, EG & G, Salem, MA) connected
to an amplifier (HCA, 28dB, DC-325 MHz; FEMTO, Berlin,)
and a digital storage oscilloscope DSA 602A with plug-in 11A52
(DC-20 MHz, Tektronix, Les Ulis, France). The kinetic data
were fitted into a multiexponential decay with a Marquart
least-square algorithm. The program was kindly provided by P.
Setif (Saclay). The very-fast phase with a t
1/2
⫽⬇3nswas
assigned to exited states of loosely bound Chls and were not
considered in the fittings.
GEPASI 3 was downloaded from http:兾兾gepasi.dbs.aber.ac.uk
softwgepasi.html (for review, see ref. 39). The simulations were
based on Scheme 1. The rate constant k
Rec
800 s
1
was taken
from the literature (33). No direct measurement of k
TyrZ
back
is
available in the literature. k
TyrZ
back
was calculated based on the
equilibrium between TyrZ
and P680
, which was estimated to
be about 10
3
at pH 8.5 (23, 40, 41). This value yields a k
TyrZ
back
of
about 3.4 10
3
s
1
based on a k
TyrZ
3.4 10
6
s
1
at pH 8.5,
as measured in the present study.
Results
Fig. 1 A, C, and E shows the EPR experiments at pH 8.5
performed with the D1-Y161F PS II mutant (i.e., the TyrZ-less
mutant), in which the typical tyrosine radical spectrum arises
solely from TyrD. Fig. 1 B, D, and F show the results obtained
with wild-type PSII.
Dark incubation of PSII from wild-type and D1-Y161F mu-
tant from Synechocystis sp. for2hatpH8.5inabuffercontaining
K
3
Fe
3
(CN)
6
and K
4
Fe
2
(CN)
6
(see Materials and Methods)
resulted in the reduction of TyrD. Fig. 1 A and B (traces labeled
‘‘dark’’) shows that in both samples, the amplitude of the TyrD
signal is less than 5% of that measured after 50 flashes (Fig. 1
A and B, traces labeled ‘‘after 50 fl’’). A further series of 50
flashes did not result in an additional increase (data not shown),
indicating that TyrD was fully oxidized after 50 flashes.
The kinetics of TyrD
andor TyrZ
were measured by mon-
itoring the EPR signal at 3,465 G (1 G 0.1 mT; see Fig. 1 A
and B arrow) in these samples. This magnetic field value is
outside the magnetic-field range in which signals from Chl and
carotenoid cations could contribute significantly (42, 43).
Fig. 1C shows the time course of TyrD
formation and decay
during a series of five flashes separated by 8 s each (flashes are
indicated by arrows) in D1-Y161F PSII. The first flash induced
a signal with an amplitude similar to that measured after 50
flashes (Fig. 1A). This observation indicates that TyrD is
transiently fully oxidized after one flash. The slow time-
resolution of the EPR spectrometer (i.e., 82 ms) used did not
allow the kinetics of TyrD
to be resolved. About 30% of the
signal decayed within 1 s, whereas the other 70% was relatively
stable. The second flash oxidized the re-reduced fraction of
TyrD, and again, about one third of this fraction decayed within
1 s. With each flash, the stable signal approached the level of
fully oxidized TyrD
(Fig. 1A). After the fifth flash, the signal
was monitored for 42 s (Fig. 1C shows only 22 s), showing the
beginning of a very slow decay (t
1/2
estimated to be 30 min).
Immediately after the recording of this trace, the TyrD
spec-
trum was measured (Fig. 1A, trace ‘‘after 5 fl’’), confirming that
almost all TyrD was oxidized after five flashes.
Scheme 1.
Faller et al. PNAS
December 4, 2001
vol. 98
no. 25
14369
BIOPHYSICS
Fig. 1E corresponds to the light-induced signal (averaged over
32 flashes) measured immediately after the recording of the
trace in C. The absence of a reversible Tyr
signal is as expected,
because TyrZ is absent in this mutant.
Fig. 1D shows the time course of Tyr
formation (TyrD
andor TyrZ
) and decay after five flashes separated by 8 s
each (flashes indicated by arrows) in wild-type PSII. The first
flash oxidized one equivalent of Tyr. About 15% of Tyr
decayed within a few seconds; the other 85% were relatively
stable. On the following flashes, one equivalent of Tyr
is
formed,
but the fraction of the signal that was stable became
smaller from flash to flash. The stable signal is attributed to
TyrD
because of the similar spectrum and decay rate as seen
for TyrD
in the Z-less mutant (Fig. 1 A and B; trace ‘‘after 5
fl’’) (see also ref. 14). Thus, after the first flash, about 85% of
the TyrD was oxidized
, while the remainder was oxidized by
the subsequent flashes.
Fig. 1F shows the flash-induced TyrZ
signal of wild-type
PSII, in which TyrD is fully oxidized after 50 preflashes (see also
ref. 44). Because it could be averaged (16 flashes, spaced by 8s),
this trace was measured with a higher time resolution and shows
that all of the TyrZ was oxidized upon each flash (i.e., it has the
same intensity as the fully oxidized TyrD in Fig. 1A).
Fig. 2 shows the flash-induced absorption changes at 820 nm
of wild-type and D1-Y161F PSII mutant from Synechocystis sp.
at pH 8.5. Chl cations absorb at this wavelength; thus, all of the
kinetics show an immediate increase caused by the fast forma-
tion of P680
followed by a decay that reflects the re-reduction
of P680
by electron donor(s).
Before the first flash, TyrD was reduced (see above) and thus
available as an electron donor to P680
. For the time window
shown in Fig. 2, the kinetic traces could be satisfactorily fitted
with three components. The initial spike present during the first
10 ns was omitted for the fits (see Material and Methods). For
the first flash on the wild-type, the fit yielded a fast phase with
t
1/2
169 ns (70%), a slower phase with t
1/2
895 ns (21%),
and a constant (9%). The constant reflects the slower decays,
which were also measured here by using a longer time range
(data not shown) and were fitted with t
1/2
112
s (4%), 820
s
(5%), and 4.3 ms (1%) (see also refs. 30 and 31). The kinetics
after the tenth flash of the wild-type were not significantly
different from the first flash [t
1/2
186 ns (67%), 950 ns (22%),
constant (11%)] (Fig. 2).
Surprisingly, the kinetic traces of the D1-Y161F mutant upon
the first flash revealed about the same t
1/2
values as the wild-
type, although the amplitudes were somewhat different [t
1/2
183 ns (45%), 675 ns (22%), constant (33%)]. The trace after the
tenth flash was very different; i.e., the constant (
sms phases)
increased dramatically at the expense of the sub-
s phases [fit:
t
1/2
222 ns (7%), 1160 ns (13%), constant (80%)]. All of the
fits of the wild-type and mutant samples at different pHs
revealed similar rates, which were in the following range: fast
phase, t
1/2
190 35 ns; slow phase, t
1/2
900 300 ns. EPR
experiments showed that almost all TyrD is oxidized after the
first flash (see above). Therefore, the fast phase that occurs upon
In Fig. 1F, TyrZ
is detected in almost all of the PSII centers. The apparent smaller amplitude
of the reversible signal in Fig. 1D is only due to a lower time resolution.
We obtained similar values for the yield of TyrD oxidation after one ash at pH 8.5 for PSII
from Synechococcus elongatus (i.e., 85%) and spinach (i.e., 55%) (data not shown). For
chloroplasts from spinach, values from 6585% were reported (34, 35).
Fig. 1. Room temperature EPR studies of TyrZ
TyrD
kinetics on Mn-
depleted PSII particles from Synechocystis sp. strain PCC 6803 at pH 8.5. The
measurements were performed on the mutant D1-Y161F (A, C, E), where
the electron donor Tyr
Z
was replaced by a Phe, and on the wild-type PSII
(B, D, F). (A and B) Traces ‘‘dark’’ show the spectrum of the dark-adapted
samples. Trace ‘‘after 5 fl’’ is the spectrum measured after 5 saturating
ashes (immediately after the experiment shown in C and D), whereas the
trace ‘‘after 50 fl’’ shows the spectrum taken after 50 ashes (four scans per
spectrum). (C and D) The time course of the EPR signal from (oxidized) Tyr
measured at 3,464 G (arrows A and B). The ve arrows indicate the
saturating laser ashes given to the samples. (E and F) The time course of
the EPR signal from (oxidized) Tyr
measured at 3,464 G (arrows in A and B)
after illumination with 50 ashes. The signal was averaged over 16 or 32
ashes (repetition rate 8 s). EPR conditions: power, 20 mW; microwave
frequency, 9.6 GHz; modulation amplitude, 4.5 G; modulation frequency,
100 kHz. Conversion time: 82 ms for CD, 5.12 ms for EF. Time constant: 41
ms for CD, 2.56 ms for EF.
Fig. 2. Flash-induced absorption transients at 820 nm in PSII particles from
Synechocystis sp. strain PCC 6803 at pH 8.5. The traces from the wild-type PSII
are denoted wt; the traces from the TyrZ-less mutant are denoted D1-Y161F.
The traces show the rst and the tenth ash after dark adaptation of the
samples with a mixture of 8 mM K
4
[Fe
2
(CN
)
6
]and8mMK
3
[Fe
3
(CN
)
6
]to
reduce TyrD. The initial amplitudes of the traces from the wild-type were
normalized to that of the D1-Y161F.
14370
www.pnas.orgcgidoi10.1073pnas.251382598 Faller et al.
the first flash, which is almost absent after the tenth flash, can
be attributed to electron donation from TyrD to P680
, with a
t
1/2
⫽⬇190 ns. The phase with a t
1/2
⫽⬇900 ns is also diminished
after the tenth flash, suggesting its assignment to P680
reduc-
tion by TyrD in a fraction of centers. The higher amplitude of the
sms phases (i.e., the constant in our fit) in the D1-Y161F
mutant than in the wild-type that occurs upon the first flash is
partially due to the presence of some residual oxidized TyrD
(5% measured by EPR) before the flash experiment (see Fig.
1 A and B and Discussion).
Treatment of the D1-Y161F PSII mutant and wild-type PSII
with NH
2
OH [instead of K
3
Fe
3
(CN)
6
K
4
Fe
2
(CN)
6
] to reduce
TyrD
did not significantly change the rates and amplitudes of
the traces, when taking into account the longer lifetime of Q
A
in
the presence of NH
2
OH.
The P680
reduction in the TyrZ-less mutant (D1-Y161F) was
also measured at different pH values (6.59.5; Fig. 3B). TyrD was
prereduced by dark-incubation in the presence of 1 mM NH
2
OH
(as verified by the lack of a TyrD
signal in EPR). The amplitude
of the long-lived P680
[i.e., the constant at this time scale (04
s); assigned to P680
Q
A
recombination] increased at the
expense of the sub-
s phases (assigned to electron donation
from TyrD) upon lowering the pH. At pH 6.5, fitting suggested
re-reduction of P680
by TyrD in 10% and by recombination
with Q
A
in 90% of the centers. This estimate is in line with the
corresponding EPR experiments, which revealed that only about
16% of all TyrD is oxidized after one flash at pH 6.5 (Fig. 3A).
Fig. 3C shows the sum of the amplitudes of the two sub-
s
phases (corresponding to electron donation from TyrD to
P680
) plotted against the pH. The experimental data show a
strong pH dependence and were fitted to a single proton
titration, which led to a pK
a
of 7.7 (Fig. 3C).
Discussion
The use of a D1-Y161F (i.e., TyrZ-less) mutant allowed a direct
measurement of the re-reduction rate of P680
by TyrD. Two
rates with t
1/2
⫽⬇190 ns and 900 ns could be attributed to this
electron donation. These rates are four to five orders of
magnitude faster than the generally accepted rate (reviewed
in ref. 45).
Buser et al. (33) reported a TyrD donation rate to P680
slower than that reported here (t
1/2
1020 ms at pH 5.07.5).
The two measurements agree at low pH (6.5) but disagree at
higher pH. Buser et al. (33) measured low yields of TyrD
at pH
7.5 in the presence of 0.4 mM ascorbate and 20 s after the flash,
and this led them to calculate a much slower rate. Experiments
(data not shown) performed under these conditions with similar
materials (PSII membranes from spinach) gave similar yields.
We suggest that the low yield is due to the ascorbate that may
reduce TyrD
after the flash and Q
A
before the flash, leading to
an underestimation of TyrD
yield.
The electron transfer rate from TyrD is very similar to that
from TyrZ to P680
at higher pH [measured in the presence
of oxidized TyrD (Fig. 2 and ref. 30)]. The sub-
s electron
donation from TyrD to P680
was measured in a significant
percentage of centers at pH 8.5, i.e., 67%. In the remaining
33% of the centers, recombination between Q
A
and P680
is
proposed to occur. In wild-type PSII with preoxidized TyrD,
the rate assigned to the recombination between Q
A
and P680
occurred in only 12% of the centers. This difference in the
amplitude of the recombination phase between wild-type and
mutant can in part be ascribed to the presence of some TyrD
(5%) in the TyrZ-less mutant before the first actinic flash
(Fig. 1 A and B, ‘‘dark’’). The reason for the remaining 16%
difference is not known.**
The amplitude of the sub-
s electron donation from TyrD to
P680
is pH dependent, with an apparent pK
a
close to pH 7.7
(Fig. 3C). At low pH, the TyrD electron donation rate is slower
than the Q
A
P680
recombination (t
1/2
0.15 ms). This pK
a
could represent the pK
a
value of a single amino acid residue. The
D2-His-189 residue is a candidate for the protonable group.
Protonation of the D2-His-189 nitrogen proximal to TyrD would
be expected to have drastic effect on TyrD electron donation. In
this case, at low pH, TyrD would have to donate its electron to
P680
either, without releasing its proton, which is energetically
very unfavorable (e.g., ref. 46) or with kinetic limitations due to
deprotonation (and reorientation) of the D2-His-189. However,
this explanation is in contradiction with Fourier transform
infrared (FTIR) measurements, suggesting that D2-His-189 is
deprotonated at pH 6.0 (in the presence of formate and phos-
phate) (27).
Another possibility, which is in line with the FTIR measure-
ments, is that the protonable group is TyrD itself and that its
protonation state is mediated by the protonation state of D2-
His-189 at its distal nitrogen. The faster electron-donation rate
of TyrD to P680
at higher pH (pK
a
of 7.7) is due to the
deprotonation of TyrD before oxidation, i.e., the proton is
already on the proximal nitrogen of D2-His-189 when the distal
nitrogen is deprotonated. At low pH the electron transfer is slow,
the TyrD is neutral and H-bonded to the proximal nitrogen of
D2-His-189, while the distal nitrogen of the D2-His-189 is
protonated. Thus, oxidation of TyrD is limited by H
transfer to
the D2-His-189, which in turn is limited by the release of H
from its distal nitrogen (see below).
It is known that TyrZ electron donation also shows a pH
dependence with a pK
a
between 7.0 and 8.3 (23, 24, 31), similar
to that for TyrD reported here. As for TyrD, titration of the TyrZ
itself or the neighboring D1-His-190 were suggested reasons for
the pH dependence (23, 24, 31). At high pH, the donation rates
of TyrZ and TyrD to P680
seem to be similar (190 ns, see
above). In contrast, at low pH, the rates of the dominating kinetic
**Two possible explanations were tested: (i) Working in total darkness and blocking the
measuring light beam until 2.5 ms before the measurement did not change the results.
Thus it is unlikely that TyrD was oxidized to a signicant extent by the measuring andor
stay light before the measurement occurred (ii) The incubation time (20 h instead of 2 h)
at pH 8.5 before the measurement did not affect the result, indicating the TyrD site was
pH equilibrated.
Fig. 3. (A) Room temperature EPR studies of TyrD
kinetics on Mn-depleted
PSII TyrZ-less mutant from Synechocystis sp. at pH 6.5 and pH 8.5. The kinetics
were measured at 3,464 G, and a ash was given at time 0. Other parameters
were as in Fig. 1D.(B) Flash-induced absorption transients at 820 nm in the
Mn-depleted TyrZ-less mutant (D1-Y161F) from Synechocystis sp. at different
values of pH as indicated in the gure. TyrD was prereduced by incubating for
90 min in 1 mM NH
2
OH. (C) The relative amplitude of the sum of the fast
components (1
s, corresponding to TyrD electron donation) at the different
pH values from Fig. 3B. The solid line shows a single proton titration tofthe
data points.
Faller et al. PNAS
December 4, 2001
vol. 98
no. 25
14371
BIOPHYSICS
phases are very different, with a t
1/2
1030
s for TyrZ (23,
24, 31) and a t
1/2
0.20.9 ms for TyrD. As suggested above, the
deprotonation of the neighboring His at its distal nitrogen could
be linked to the protonation of its proximal nitrogen by the
proton from the Tyr. This process could be easier with TyrZ than
with TyrD, because D1-Glu-189 appears to be (indirectly)
associated with a TyrZ deprotonation pathway (47), possibly
positioning the proton acceptor for D1-His-190. No equivalent
carboxylic acid is conserved in the D2 sequence associated with
TyrD, where it is replaced by the acidbase inactive Phe residue
(4, 46, 4851). Another explanation for TyrZ oxidation being
faster than TyrD oxidation at low pH could be the possibility that
TyrZ
has multiple H-bonding partners (for review, see ref. 3;
refs. 25, 26, 52), whereas TyrD
is considered to have D2-His-189
as its unique H-bonding partner (19).
Whereas in the TyrZ-less PSII mutant (D1-Y161F), TyrD is
oxidized directly by P680
(see above), in the wild-type PSII,
TyrD could be oxidized in two ways: (i) directly by P680
, and
(ii) by rapid formation of a TyrZP680
º TyrZ
(H
)P680
equilibrium followed by a slower oxidation of TyrD by the
equilibrium population of P680
. Because of the lower redox-
potential of TyrD (18, 34, 53), it would act as a thermodynamic
sink for the electron hole (see Scheme 1). The second of these
mechanisms has been the generally accepted view. Our obser-
vation of a rapid oxidation of TyrD by P680
(t
1/2
⫽⬇190 ns)
in the D1-Y161F mutant strongly suggests that a direct rapid
route also exists in wild type, but it does not rule out the other
mechanism.
In this context, an observation of Boussac and Etienne (34) is
relevant. They showed in Tris-washed thylakoids with prere-
duced TyrD that the yield of TyrD
after one flash was almost
the same (65%) in the absence and in the presence of
phenylenediamine, a rapid electron donor to TyrZ
. We have
performed similar experiments by using Mn-depleted PSII mem-
branes from spinach and confirmed these results (data not
shown). Because phenylenediamine donates to TyrZ
withat
1/2
1 ms (34, 54), both electron donation rates from TyrD to
P680
and from P680
to TyrZ
(i.e., k
TyrD
and k
TyrZ
back
) must
be substantially faster than this; otherwise, the yield of TyrD
would decrease dramatically in the presence of phenlyenedia-
mine. Simulation of Scheme 1 using the program
GEPASI (see
Materials and Methods) in the presence of phenylenediamine
(i.e., k
Don
700 s
1
instead of 4 s
1
) resulted in a formation of
65% TyrD
only when k
TyrD
was larger than 1 10
6
s
1
(i.e., t
1/2
700 ns). Formation of 85% TyrD
needs a k
TyrD
3.7 10
6
(i.e., t
1/2
⫽⬇190 ns). These simulations provide an argument
that TyrD in wild-type PSII is oxidized in the sub-
s range at pH
8.5, and that in a fraction of centers, TyrD is oxidized directly by
P680
(mechanism 1) with a similar rate as in the D1-Y161F
mutant.
Before the present work, the reports that electron donation
from TyrD to P680
were slow had been rationalized by
assuming a high reorganization energy caused by the trapping of
the departing proton from TyrD by D2-His-189 (4). Thus, the
more rapid donation from TyrZ to P680
was taken as indicating
that no such charge was accumulated close to TyrZ
. This line of
thought in turn was used as an argument in favor of the
translocation of the proton away from TyrZ
, a feature that could
be taken as being in favor of certain aspects of the H-atom
(electron proton) abstraction model for oxygen evolution (46).
The present work questions the need to invoke such a large
reorganization energy associated with the putative charge close
to TyrD
and thus weakens this argument for proton transloca-
tion away from the TyrZ
site.
The reassessment of the relative donation rates of TyrZ and
TyrD to P680
has repercussions on the question of the local-
ization of P680
(for instance, see ref. 45). There are four Chls
in the inner reaction center symmetrically positioned around a
C
2
axis (ref. 6; Protein Databank accession no. 1FE1). It was
suggested that the much faster donation rate of TyrZ compared
with TyrD could be explained if P680
were located on the D1
side (i.e., closer to TyrZ than TyrD) of the reaction center (for
review, see ref. 45). The similar donation rates found here
question this argument. According to Page et al. (55), who
describe a correlation between the distance and the rate of an
electron transfer in proteins, an electron donation rate of t
1/2
190 ns gives an upper limit of 14 Å for the edge-to-edge
distance between the donor and acceptor. In the structural
model of PSII, only the Chls on the D2 side (P
D2
and Chl
D2
) are
located within this distance from TyrD (the distance P
D1
to TyrD
is 18 Å edge to edge), and thus a direct electron transfer from
TyrD to P
D1
is unlikely to occur with a t
1/2
190 ns. This
argument suggests that P680
is at least partially localized on the
Chl P
D2
(located on the D2 side). A recent study, performed in
the presence of TyrD
, suggested that the cation is predominantly
localized on P
D1
(D1 side; ref. 12). These results are not
contradictory, because the present studies suggest only partial
localization of P680
on P
D2
and only in the presence of reduced
TyrD (TyrD was oxidized in ref. 12). An effect of the oxidation
state of TyrD on the localization of P680
is conceivable (see
below).
In the present work, the measured electron-transfer rates to
P680
from TyrD and TyrZ were the same in Mn-depleted
PSII at pH 8.5 (t
1/2
190 ns for the dominating phase). Thus,
when both TyrD and TyrZ are able to donate electrons to
P680
(e.g., first flash in wild type), the reduction rate P680
is expected to be twice as fast, i.e., t
1/2
95 ns. The measured
rate, however, exhibited again a t
1/2
190 ns (Fig. 2, ‘‘wt:1.
flash’’). The most straightforward explanation for this finding
is that the redox-state of TyrD affects the TyrZ donation. The
following assumptions would allow for such a situation: (i) the
TyrZ donation rate is slower than t
1/2
190 ns when both
tyrosines are reduced. This assumption seems reasonable given
the environment and potential of the two tyrosines (see above)
(ii) The presence of TyrD
accelerates the TyrZ to P680
electron donation. It seems reasonable that the positive charge
formed close to the TyrD site upon its oxidation (for example,
the protonated D2-His-189) has an electrostatic effect on
P680
(55), increasing its potential and perhaps shifting the
location of the positive charge (P680
) over to the D1 side
(P
D1
) (see ref. 56) where it is now supposed to be (12). A
similar role was suggested for D2-Arg-181 (57). Such a role for
TyrD
on the redox-potential of P680
, and thus on the kinetics
of TyrZ oxidation, could provide a mechanistic raison deˆtre
for TyrD in addition to its postulated redox role in photoac-
tivation and oxidation of the lower S states (17). Experimental
indications of an interaction of TyrD and P680 have been
reported: (i) TyrD oxidation induces an electrochromic shift in
the absorption spectrum of the P680 Soret band (58); (ii)
charge recombination between Q
A
and TyrZ
was slower in
wild-type PSII (in the presence of TyrD
) than in TyrD-less
mutants (59). These findings can be explained by an electro-
static effect of TyrD
H
-His on P680
that would raise the
redox potential of P680
and thus depopulate P680
(shift the
equilibrium of TyrZ
º P680
to the left). This depopulation
of P680
, which is an intermediate in the Q
A
TyrZ
recom-
bination (45), would slow down the recombination, as ob-
served. However, Hays et al. (28) compared the TyrZ donation
in wild-type (TyrD
present) and D2-Y161F mutant at differ-
ent pHs (7.09.0) and found no significant differences, indi-
cating that the presence of TyrD
does not affect the TyrZ
donation. Further comparative studies on the TyrD-less and
TyrZ-less mutant, as well as on the wild-type, should clarify
whether the redox chemistry of TyrD has an electrostatic
effect on P680
.
14372
www.pnas.orgcgidoi10.1073pnas.251382598 Faller et al.
We thank Dr. P. Se´tif for providing the software for the kinetic analyses;
A. P. Nguyen for isolating the thylakoid membranes from and maintaining
the wild-type and mutant cultures of Synechocystis 6803; Drs. F. Rappaport,
S. Un, T. Mattioli, A. Ivancich, C. Berthomieu, and R. Hienerwadel for
helpful discussions; Drs. B. Diner, E. Schlodder, P. Nixon, W. Coleman, F.
Rappaport, J. Lavergne, W. Vermaas, and D. Chisholm for providing us
with a preprint of ref. 12. The research was supported by a grant from the
Swiss National Science Foundation (to P.F.), a European Union Training
and Mobility of Researchers (TMR) Research Network grant (FMRX-
CT982014), a cooperation grant ‘‘metalloproteins Japan’’ from the Min-
istre des Affaires E
´
trangers (PO N°20001577), and National Institutes of
Health Grant GM43496 (to R.J.D.).
1. Stubbe J. & van der Donk, W. A. (1998) Chem. Rev. (Washington, D.C.) 98,
705762.
2. Diner, B. A. (2001) Biochim. Biophys. Acta 1503, 147163.
3. Debus, R. J. (2001) Biochim. Biophys. Acta 1503, 164186.
4. Babcock, G. T., Espe, M., Hoganson, C., Lydakis-Simantriris, N., McCracken,
J., Shi, W., Styring, S., Tommos, C. & Warncke, K. (1997) Acta Chem. Scand.
51, 533540.
5. Rappaport, F. & Lavergne, J. (2001) Biochim. Biophys. Acta 1503, 246259.
6. Zouni, A., Witt, H.-T., Kern, J., Fromme, P., Krauss, N., Saenger, W. & Orth,
P. (2001) Nature (London) 409, 739743.
7. Debus, R. J., Barry, B. A., Babcock, J. T. & McIntosh, L. (1988) Proc. Natl.
Acad. Sci. USA 85, 427430.
8. Vermass, W. F. J., Rutherford, A. W. & Hansson, O. (1988) Proc. Natl. Acad.
Sci. USA 85, 84778481.
9. Hirsh, D. J. & Brudvig, G. W. (1993) J. Phys. Chem. 97, 1321613222.
10. Koulougliotis, D., Tang, X.-S., Diner, B. A.&&Brudvig,G.W.(1995)
Biochemistry 34, 28502856.
11. Astashkin, A. V., Kodera, Y. & Kawamori, A. (1994) Biochim. Biophys. Acta
1187, 8993.
12. Diner, B. A., Schlodder, E., Nixon, P. J., Coleman, W. J., Rappaport, F.,
Lavergne, J., Vermaas, W. F. J. & Chisholm, D. A. (2001) Biochemistry 40,
92659281.
13. Rutherford, A. W. & Nitschke W. (1996) in Origin and Evolution of Biological
Energy Conversion, ed. Baltscheffsky, H. (VCH, New York), pp. 14375.
14. Babcock, G. T. & Sauer, K. (1975) Biochim. Biophys. Acta 376, 315328.
15. Brettel, K., Schlodder, E. & Witt, H. T. (1984) Biochim. Biophys. Acta 766,
403415.
16. Babcock, G. T. & Sauer, K. (1973) Biochim. Biophys. Acta 325, 483503.
17. Styring, S. & Rutherford, A. W. (1987) Biochemistry 26, 24012405.
18. Vass, I. & Styring, S. (1991) Biochemistry 30, 830839.
19. Campell, K. A., Peloquin, J. M., Diner, B. A., Tang, X. S., Chisholm, D. A. &
Britt, R. D. (1997) J. Am. Chem. Soc. 119, 47874788.
20. Rodriguez, I. D., Chandrashekar, T. K. & Babcock, G. T. (1987) in Progress in
Photosynthesis Research, ed. Biggins, J. (Martinus Nijhoff, Dordrecht, The
Netherlands), Vol. I, pp. 471474.
21. Tang, X.-S., Chisholm, D. A., Dismukes, G. C., Brudvig, G. W. & Diner, B. A.
(1993) Biochemistry 32, 1374213748.
22. Tommos, C., McCracken, J., Styring, S. & Babcock, G. T. (1998) J. Am. Chem.
Soc. 120, 1044110452.
23. Diner, B. A., Force, D. A., Randall, D. W. & Britt, R. D. (1998) Biochemistry
37, 1793117943.
24. Ahlbrink, R., Haumann, M., Cherapanov, D., Bogenshausen, O., Mulkidjanian,
A. & Junge W. (1998) Biochemistry 37, 11311142.
25. Force, D. A., Randall, D. W., Britt, R. D., Tang, X. S. & Diner, B. A. (1995)
J. Am. Chem. Soc. 117, 1054710554.
26. Tommos, C., Tang, X. S., Warncke, K., Hoganson, C. W., Styring, S.,
McCracken, J. & Diner, B. A. (1995) J. Am. Chem. Soc. 117, 1032510335.
27. Hienerwadel, R., Boussac, A., Breton, J., Diner, B. & Berthomieu, C., (1997)
Biochemistry 36, 1471214723.
28. Berthomieu, C., Hienerwadel, R., Boussac, A., Breton, J. & Diner, B. (1998)
Biochemistry 37, 1054710554.
29. Conjeaud, H. & Mathis, P. (1980) Biochim. Biophys. Acta 590, 353359.
30. Schlodder, E. & Meyer, B. (1987) Biochim. Biophys. Acta 890, 2331.
31. Hays, A. A., Vassiliev, I. R., Golbeck, J. H. & Debus, R. J. (1999) Biochemistry
38, 1185111865.
32. Hays, A. A., Vassiliev, I. R., Golbeck, J. H. & Debus, R. J. (1998) Biochemistry
37, 1135211365.
33. Buser, C. A., Thompson, L. K., Diner, B. A. & Brudvig, G. W. (1990)
Biochemistry 29, 89778985.
34. Boussac, A. & Etienne, A. L. (1982) Biochem. Biophys. Res. Commun. 109,
12001205.
35. Boussac, A. & Etienne, A. L. (1982) FEBS Lett. 148, 113116.
36. Debus, R. J., Barry, B. A., Sithole, I., Babcock, G. T. & McIntosh, L. (1988)
Biochemistry 27, 90719074.
37. Debus, R. J., Nguyen, A. P. & Conway, A. B. (1990) in Current Research in
Photosynthesis, ed. Baltscheffsky, M. (Kluwer, Dordrecht, The Netherlands),
Vol. I, pp. 829832.
38. Debus, R. J., Campbell, K. A., Gregor, W., Li, Z.-L., Burnap, R. L. & Britt,
R. D. (2001) Biochemistry 40, 36903699.
39. Mendes, P. (1997) Trends Biochem. Sci. 22, 361363.
40. Yerkes, C. T., Babcock, G. T. & Crofts, A. R. (1983) Biochim. Biophys. Acta
158, 359363.
41. Rappaport, F. & Lavergne, J. (1997) Biochemistry 36, 1529415303.
42. Visser, J. W. M., Rijgersberg, C. P. & Gast, P. (1977) Biochim. Biophys. Acta
460, 3646.
43. Hanley, J., Deligiannakis, Y., Pascal, A., Faller, P. & Rutherford, A. W. (1999)
Biochemistry 38, 81898185.
44. Blankenship, R. E., Babcock, G. T., Warden, J. T. & Sauer, K. (1975) FEBS
Lett. 51, 287293.
45. Diner, B. A. & Babcock, G. T. (1996) in Oxygenic Photosynthesis: The Light
Reactions, eds. Ort, D. R. & Yocum, C. F. (Kluwer, Dordrecht), pp. 213247.
46. Tommos, C. & Babcock, G. T. (2000) Biochim. Biophys. Acta 1458, 199219.
47. Debus, R. J., Campbell, K.A., Pham, D. P., Hays, A.-M. A. & Britt, R. D. (2000)
Biochemistry 39, 6276287.
48. Svensson, B., Vass, I., Cedergren, E. & Styring, S. (1990) EMBO J. 9,
20512059.
49. Ruffle, S., Donnelly, D., Blundell, T. L. & Nugent, J. H. A. (1992) Photosynth.
Res. 34, 287300.
50. Chu, H. A., Nguyen, A. P. & Debus, R. J. (1995) Biochemistry 34, 58395858.
51. Svensson, B., Etchebest, C., Tuffery, P., van Kan, P., Smith, J. & Styring, S.
(1996) Biochemistry 35, 1448614502.
52. Un, S., Tang, X. S. & Diner, B. A. (1996) Biochemistry 35, 679684.
53. Boussac, A. & Etienne, A. L. (1984) Biochim. Biophys. Acta 766, 576584.
54. Diner, B. A. (1977) in Proceedings of the 4
th
International Congress on
Photosynthesis, ed. Hall, D. O., Coombs, J. & Goodwin, T. W. (Biochemical
Society, London), pp. 359372.
55. Page, C. C., Moser, C. C., Chen, X. & Dutton, P. L. (1999) Nature (London)
402, 4752.
56. Nugent, J. H. A., Bratt, P. J., Evans, M. C. W., MacLachlan, D. J., Rigby, S. E. J.,
Ruffle, S. V. & Turconi, S. (1994) Biochem. Soc. Trans. 22, 327331.
57. Mulkidjanian, A. (1999) Biochim. Biophys. Acta 1410, 16.
58. Diner, B., Tang, X-S., Zheng, M., Dismukes, G. C., AnnForce, D., Randall,
D. W. & Britt, R. D. (1995) in Photosynthesis: From Light to Biosphere, ed.
Mathis, P. (Kluwer, Dordrecht, The Netherlands) pp. 229234.
59. Boerner, R. J., Bixby, K. A., Nguyen, A. P., Noren, G. H., Debus, R. J. & Barry,
B. A. (1993) J. Biol. Chem. 268, 18171823.
Faller et al. PNAS
December 4, 2001
vol. 98
no. 25
14373
BIOPHYSICS
... In Mn-depleted PSII, the electron transfer from Tyr Z to [P D1 P D2 ] + is strongly pH dependent (Conjeaud and Mathis 1980) with a t 1/2 of ~ 190 ns at pH 8.5 and ~ 2-10 µs at pH 6.5 (Faller et al. 2001). At pH 8.5 and above, Tyr D , the Tyr160 of the D2 polypeptide that is symmetrically positioned with respect to Tyr Z (Fig. 1), is slowly reduced in the dark (Boussac and Etienne 1982) and becomes able to donate an electron to P D1 + with a t 1/2 of ~ 190 ns, similar to Tyr Z donation (Faller et al. 2001). ...
... In Mn-depleted PSII, the electron transfer from Tyr Z to [P D1 P D2 ] + is strongly pH dependent (Conjeaud and Mathis 1980) with a t 1/2 of ~ 190 ns at pH 8.5 and ~ 2-10 µs at pH 6.5 (Faller et al. 2001). At pH 8.5 and above, Tyr D , the Tyr160 of the D2 polypeptide that is symmetrically positioned with respect to Tyr Z (Fig. 1), is slowly reduced in the dark (Boussac and Etienne 1982) and becomes able to donate an electron to P D1 + with a t 1/2 of ~ 190 ns, similar to Tyr Z donation (Faller et al. 2001). Consequently, at room temperature and pH 8.5, the reduced Tyr D can donate an electron to [P D1 P D2 ] + in approximately half of the centers upon a saturating ns flash in Mn-depleted PSII, while in the remaining half of the centers [P D1 P D2 ] + is reduced by Tyr Z . ...
... The material was a Mn-depleted PsbA3/PSII which was dark adapted for ~ 3-4 h at pH 8.6. This long dark-incubation allows Tyr D to be reduced in the great majority of the centers (Boussac and Etienne 1982;Faller et al. 2001). The black spectrum was recorded after the first flash, i.e., it corresponds to the formation of the [P D1 P D2 ] + Q A − state when Tyr D is not yet oxidized. ...
Article
Full-text available
Flash-induced absorption changes in the Soret region arising from the [PD1PD2]+ state, the chlorophyll cation radical formed upon light excitation of Photosystem II (PSII), were measured in Mn-depleted PSII cores at pH 8.6. Under these conditions, TyrD is i) reduced before the first flash, and ii) oxidized before subsequent flashes. In wild-type PSII, when TyrD● is present, an additional signal in the [PD1PD2]+-minus-[PD1PD2] difference spectrum was observed when compared to the first flash when TyrD is not oxidized. The additional feature was “W-shaped” with troughs at 434 nm and 446 nm. This feature was absent when TyrD was reduced, but was present (i) when TyrD was physically absent (and replaced by phenylalanine) or (ii) when its H-bonding histidine (D2-His189) was physically absent (replaced by a Leucine). Thus, the simple difference spectrum without the double trough feature at 434 nm and 446 nm, seemed to require the native structural environment around the reduced TyrD and its H bonding partners to be present. We found no evidence of involvement of PD1, ChlD1, PheD1, PheD2, TyrZ, and the Cytb559 heme in the W-shaped difference spectrum. However, the use of a mutant of the PD2 axial His ligand, the D2-His197Ala, shows that the PD2 environment seems involved in the formation of “W-shaped” signal.
... In Mn-depleted PSII, the electron transfer from TyrZ to [PD1PD2] + is strongly pH dependent (Conjeaud and Mathis 1980) with a t1/2 of ∼ 190 ns at pH 8.5 and ∼ 2-10 µs at pH 6.5 (Faller et al. 2001). At pH 8.5 and above, TyrD, the Tyr160 of the D2 polypeptide that is symmetrically positioned with respect to TyrZ (Fig. 1), is slowly reduced in the dark (Boussac and Etienne 1982) and becomes able to donate an electron to PD1 + with a t1/2 of ∼ 190 ns, similar to TyrZ donation (Faller et al. 2001). ...
... In Mn-depleted PSII, the electron transfer from TyrZ to [PD1PD2] + is strongly pH dependent (Conjeaud and Mathis 1980) with a t1/2 of ∼ 190 ns at pH 8.5 and ∼ 2-10 µs at pH 6.5 (Faller et al. 2001). At pH 8.5 and above, TyrD, the Tyr160 of the D2 polypeptide that is symmetrically positioned with respect to TyrZ (Fig. 1), is slowly reduced in the dark (Boussac and Etienne 1982) and becomes able to donate an electron to PD1 + with a t1/2 of ∼ 190 ns, similar to TyrZ donation (Faller et al. 2001). Consequently, at room temperature and pH 8.5, the reduced TyrD can donate an electron to [PD1PD2] + in approximately half of the centers upon a saturating ns flash in Mn-depleted PSII, while in the remaining half of the centers [PD1PD2] + is reduced by TyrZ. ...
... With the semi-logarithmic plot used in Panel B of Fig. 3, the decays at both 432 nm and 446 nm were almost linear from 10 ns to 1 µs and had a similar t1/2 of ~ 200 ns. The P680 + reduction has been studied earlier in detail in the conditions used (Faller et al. 2001, Rappaport et al. 2009). The t1/2 value in Fig. 3B is very close to the value of 190 ns found previously (Faller et al. 2001, Rappaport et al. 2009). ...
Preprint
Full-text available
Flash-induced absorption changes in the Soret region arising from the (PD1PD2)+ state, the chlorophyll cation radical formed upon light excitation of Photosystem II (PSII), were measured in Mn-depleted PSII cores at pH 8.6. Under these conditions, TyrD is i) reduced before the first flash, and ii) oxidized before subsequent flashes. In wild-type PSII, when TyrDox is present, an additional signal in the (PD1PD2)+-minus-(PD1PD2) difference spectrum was observed when compared to the first flash when TyrD is not oxidized. The additional feature was “W-shaped” with troughs at 434 nm and 446 nm. This feature was absent when TyrD was reduced, but was present i) when TyrD was physically absent (and replaced by phenylalanine) or ii) when its H-bonding histidine (D2-His189) was physically absent (replaced by a Leucine). Thus, the simple difference spectrum without the double trough feature at 434 nm and 446 nm, seemed to require the native structural environment around the reduced TyrD and its H bonding partners to be present. We found no evidence of involvement of PD1, ChlD1, PheD1, PheD2, TyrZ, and the Cytb559 heme in the W-shaped difference spectrum. However, the use of a mutant of the PD2 axial His ligand, the D2-His197Ala, shows that the PD2 environment seems involved in the formation of “W-shaped” signal.
... In Mn-depleted PSII, the electron transfer from Tyr Z to P D1 + is strongly pH dependent (Conjeaud and Mathis 1980) with a t 1/2 of ∼ 190 ns at pH 8.5 and ∼ 2-10 µs at pH 6.5 (Faller et al. 2001 The PSII from C. thermalis grown under far-red light was puri ed as previously described (Nürnberg et al. 2018), and then also treated with NH 2 OH as described above. The D2/H197A with a His-tag was puri ed with the same protocol as T. elongatus except the breaking of the cells with the French press that was done after resuspension of the cells in 100 mM Tris pH 8.0. ...
... PsbA1 vs PsbA3, does not affect the formation and spectrum of the additional "W-shape" structure observed after the 5th to 10th ashes. (Faller et al. 2001). Finally, this negative spectral feature is not observed after the rst ash, while Q A − is also formed on this rst ash, con rming that the spectral feature does not arise from Q A − itself. ...
... Clearly, the decay is ∼ 5 times slower in the D2/H197A-PSII. This slowdown is not due to a species effect (Synechocystis vs T. elongatus) because we have seen that the decay of [P D1 P D2 ] + (measured at 820 nm) occurred with the same rate (t 1/2 ∼ 200 ns) in both species (Faller et al. 2001). The possibility that the cation is more localized on P D2 in the D2/H197A is therefore a likely hypothesis. ...
Preprint
Full-text available
Flash-induced absorption changes in the Soret region arising from the [P D1 P D2 ] ⁺ state, the chlorophyll cation radical formed upon excitation of Photosystem II (PSII), were obtained using Mn-depleted PSII cores at pH 8.6. Under these conditions, Tyr D is i ) reduced before the first flash, and ii ) oxidized before subsequent flashes. In wild-type PSII, when Tyr D ● is present, an additional signal in the [P D1 P D2 ] ⁺ - minus -[P D1 P D2 ] difference spectrum was observed when compared to the first flash when Tyr D is not oxidized. The additional feature was “W-shaped” with troughs at 434 nm and 446 nm. This feature was absent when Tyr D was reduced, but was present i ) when Tyr D was physically absent (and replaced by phenylalanine) or ii ) when its H-bonding histidine (D2-His190) was physically absent (replaced by a Leucine). Thus, the simple difference spectrum without the double trough feature at 434 nm and 446 nm, seemed to require the native structural environment around the reduced Tyr D and its H bonding partners to be present. We found no evidence of involvement of P D1 , Chl D1 , Phe D1 , Phe D2 , Tyr Z , and the Cyt b 559 heme in the W-shaped difference spectrum. However, and surprisingly, the use of a mutant of the P D2 axial His ligand, the D2-His197Ala, shows that the P D2 environment seems involved in the “W-shaped” signal.
... In Mn-depleted PSII, the electron transfer from TyrZ to PD1 + is strongly pH dependent (Conjeaud and Mathis 1980) with a t1/2 of  190 ns at pH 8.5 and  2-10 µs at pH 6.5 (Faller et al. 2001). At pH 8.5, TyrD, the Tyr160 of the D2 polypeptide that is symmetrically positioned with respect to TyrZ (Fig. 1), is slowly reduced in the dark (Boussac and Etienne 1982) and becomes able to donate an electron to PD1 + with a t1/2 of  190 ns, similar to TyrZ donation (Faller et al. 2001). ...
... In Mn-depleted PSII, the electron transfer from TyrZ to PD1 + is strongly pH dependent (Conjeaud and Mathis 1980) with a t1/2 of  190 ns at pH 8.5 and  2-10 µs at pH 6.5 (Faller et al. 2001). At pH 8.5, TyrD, the Tyr160 of the D2 polypeptide that is symmetrically positioned with respect to TyrZ (Fig. 1), is slowly reduced in the dark (Boussac and Etienne 1982) and becomes able to donate an electron to PD1 + with a t1/2 of  190 ns, similar to TyrZ donation (Faller et al. 2001). Consequently, at room temperature and pH 8.5, the reduced TyrD can donate an electron to PD1 + in approximately half of the centers upon a saturating ns flash in Mn-depleted PSII, while in the remaining half PD1 + is reduced by TyrZ. ...
... With the semi-logarithmic plot used in Panel B of Fig. 3, the decays at both 432 nm and 446 nm were almost linear from 10 ns to 1 µs and had a similar t1/2 of ~ 200 ns. This t1/2 value is very close to the value of 190 ns found previously (Faller et al. 2001, Rappaport et al. 2009). ...
Preprint
Full-text available
Flash-induced absorption changes in the Soret region arising from the [P D1 P D2 ] ⁺ state, the chlorophyll cation radical formed upon excitation of Photosystem II (PSII), were obtained using Mn-depleted PSII cores at pH 8.6. Under these conditions, Tyr D is reduced before the first flash but oxidised before subsequent flashes. When Tyr D • is present, an additional signal in the [P D1 P D2 ] ⁺ - minus -[P D1 P D2 ] difference spectrum was observed when compared to the first flash. The additional feature was W-shaped with troughs at 434 nm and 446 nm. This feature was absent when Tyr D was reduced, but was present when Tyr D was physically absent (and replaced by phenylalanine) or when its H-bonding histidine (D2-His190) was physically absent (replaced by a Leucine),. Thus, the simple difference spectrum without the double trough feature at 434 nm and 446 nm, required the native structural environment around the reduced Tyr D and its H bonding partners to be present. A range of PSII variants were surveyed, and we found no evidence of involvement of P D1 , Chl D1 , Phe D1 , Phe D2 , Tyr Z , and the Cyt b 559 heme in difference spectrum. Direct data ruling out the participation of P D2 is lacking. It seems possible that the specific H-bonding environment of around reduced Tyr D allows a more homogenous electrostatic environment for [P D1 P D2 ] ⁺ . A role for P D2 in the double-trough Soret signal may be tested using mutants of P D2 axial His ligand D2-His197.
... The copyright holder for this preprint this version posted February 14, 2023. ; https://doi.org/10.1101/2023.02.13.528314 doi: bioRxiv preprint that measured in Mn-depleted PSII in which the t1/2 is  10-20 µs (not shown), i.e. in the same time range as the oxidation of TyrZ by P680 +• [40]. ...
Preprint
Full-text available
In the cyanobacterium Thermosynechococcus elongatus , there are three psbA genes coding for the Photosystem II (PSII) D1 subunit that interacts with most of the main cofactors involved in the electron transfers. Recently, the 3D crystal structures of both PsbA2-PSII and PsbA3-PSII have been solved [Nakajima et al., J. Biol. Chem. 298 (2022) 102668.]. It was proposed that the loss of one hydrogen bond of Phe D1 due to the D1-Y147F exchange in PsbA2-PSII resulted in a more negative E m of Phe D1 in PsbA2-PSII when compared to PsbA3-PSII. In addition, the loss of two water molecules in the Cl-1 channel was attributed to the D1-P173M substitution in PsbA2-PSII. This exchange, by narrowing the Cl-1 proton channel, could be at the origin of a slowing down of the proton release. Here, we have continued the characterization of PsbA2-PSII by measuring the thermoluminescence from the S 2 Q A ⁻ /DCMU charge recombination and by measuring the time-resolved absorption changes of the dye bromocresol purple. It was found that i ) the E m of Phe D1 −• /Phe D1 was decreased by ∼ 30 mV in PsbA2-PSII when compared to PsbA3-PSII and ii ) the kinetics of the proton release into the bulk was significantly slowed down in PsbA2-PSII in the S 2 Tyr Z • to S 3 Tyr Z and S 3 Tyr Z • → (S 3 Tyr Z • )’ transitions. This slowing down was partially reversed by the PsbA2/M173P mutation and induced by the PsbA3/P173M mutation thus confirming the role of the D1-P173 in the function of the Cl-1 channel.
... 5 TyrD is not involved in the electron transfer pathway from the Mn 4 CaO 5 cluster. However, the neutral deprotonated radical TyrD-O • is formed upon oxidation of TyrD-OH by [P D1 /P D2 ] •+19−22 and this occurs in a tens of ms timescale 23 (also see ref 24). The proton released from TyrD-OH is transferred along the downhill proton transfer pathway that proceeds via D2-Arg180 and a series of waters toward D2-His61, the protein bulk surface. ...
Article
Full-text available
In photosystem II (PSII), the second-lowest oxidation state (S1) of the oxygen-evolving Mn4CaO5 cluster is the most stable, as the radical form of the redox-active D2-Tyr160 is considered to be a candidate that accepts an electron from the lowest oxidation state (S0) in the dark. Using quantum mechanical/molecular mechanical calculations, we investigated the redox potential (E m) of TyrD and its H-bond partner, D2-His189. The potential energy profile indicates that the release of a proton from the TyrD...D2-His189 pair leads to the formation of a low-barrier H-bond. The E m depends on the H+ position along the low-barrier H-bond, e.g., 680 mV when the H+ is at the D2-His189 moiety and 800 mV when the H+ is at the TyrD moiety, which can explain why TyrD mediates both the S0 to S1 oxidation and the S2 to S1 reduction.
Article
In the cyanobacterium Thermosynechococcus elongatus, there are three psbA genes coding for the Photosystem II (PSII) D1 subunit that interacts with most of the main cofactors involved in the electron transfers. Recently, the 3D crystal structures of both PsbA2-PSII and PsbA3-PSII have been solved [Nakajima et al., J. Biol. Chem. 298 (2022) 102668.]. It was proposed that the loss of one hydrogen bond of PheD1 due to the D1-Y147F exchange in PsbA2-PSII resulted in a more negative Em of PheD1 in PsbA2-PSII when compared to PsbA3-PSII. In addition, the loss of two water molecules in the Cl-1 channel was attributed to the D1-P173M substitution in PsbA2-PSII. This exchange, by narrowing the Cl-1 proton channel, could be at the origin of a slowing down of the proton release. Here, we have continued the characterization of PsbA2-PSII by measuring the thermoluminescence from the S2QA-/DCMU charge recombination and by measuring proton release kinetics using time-resolved absorption changes of the dye bromocresol purple. It was found that i) the Em of PheD1-•/PheD1 was decreased by ~30 mV in PsbA2-PSII when compared to PsbA3-PSII and ii) the kinetics of the proton release into the bulk was significantly slowed down in PsbA2-PSII in the S2TyrZ• to S3TyrZ and S3TyrZ• → (S3TyrZ•)' transitions. This slowing down was partially reversed by the PsbA2/M173P mutation and induced by the PsbA3/P173M mutation thus confirming a role of the D1-173 residue in the egress of protons trough the Cl-1 channel.
Article
Tyrosine radicals are notoriously short-lived/unstable in solution, while they present an impressive degree of stability and versatility in bioenzymes. Herein, we have developed a library of hybrid biomimetic materials (HBMs), which consists of tyrosine-containing oligopeptides covalently grafted on SiO2 nanoparticles, and studied the formation, lifetime, and redox properties of tyrosyl radicals. Using electron paramagnetic resonance spectroscopy, we have studied the radical-spin distribution as a probe of the local microenvironment of the tyrosyl radicals in the HBMs. We find that the lifetime of the tyrosyl radical can be enhanced by up to 6 times, by adjusting three factors, namely, a proximal histidine, the length of the oligopeptide, and the interface with the SiO2 nanomatrix. This is shown to be correlated to a significant lowering of E1/2 from +736 mV, in free tyrosine, to +548 mV in the {12-peptide}@SiO2 material. Moreover, we show that grafting on SiO2 lowers the E1/2 of tyrosine radicals by ∼50 mV in all oligopeptides. Analysis of the spin-distribution by EPR reveals that the positioning of a histidine at a H-bonding distance from the tyrosine further favors tyrosine radical stabilization.
Thesis
Bei der lichtinduzierten Oxidation von Wasser im Photosystem II (PSII) werden zwei wassermoleküle im katalytischen Zyklus des Metallclusters (Mn4CaO5) benötigt, und vier Protonen aus dem Cluster in den Lumen abgegeben. Daher ist es für das Verständnis des Mechanismus´ der Wasseroxidation von entscheidender Bedeutung, die Veränderung der Protonierungszustände am cluster während der Katalyse zu untersuchen. Hierbei sollten sowohl die Wasserkanäle für die Zuführung der Substratwassermoleküle als auch die Transportwege für die Freisetzung der Protonen untersucht werden. Deshalb wurde in meiner ersten Veröffentlichung ein neues Protokoll entwickelt, um einzelne große Kristalle von dPSII mit einer Länge von ~3 mm in der Längsachse zu züchten. Diese Kristalle mit einer Auflösung von ca. 8 Å gemessen. Um eine höhere Auflösung zu erzielen, ist die Verbesserung der Kristallqualität essenziell. Daher wurde in meiner zweiten Veröffentlichung die Struktur des Detergens-Protein-Komplexes von dPSII mit βDM, durch Anwendung von SANS in Kombination mit SAXS untersucht. Die Ergebnisse zeigten, dass βDM eine monomolekulare Schicht um dPSII bildet. Darüber hinaus konnten freie Mizellen von βDM in der Lösung nachgewiesen werden. Damit ist eine weitere Optimierung der βDM-Konzentration in der Proteinlösung erforderlich, um die Bildung von freien Mizellen zu minimieren. In meiner dritten Veröffentlichung wurde die strukturelle Dynamik in den Wasserkanälen, während des S2-S3 Übergangs mit Hilfe der XFEL untersucht. Ein Datensatz mit einer hohen Auflösung von 1,89 Å wurde durch die Zusammenführung von Daten gewonnen, die während des S2-S3 Übergangs gesammelt wurden. In Anbetracht der Analyse der zusammengeführten Daten und der einzelnen Zeitpunkte, die während des S2-S3 Übergangs gesammelt wurden, ist es wahrscheinlich, dass ein Substratwasser durch den O1-Kanal geliefert wird. Im Gegensatz dazu wird ein Proton aus dem Cluster durch den Cl1 Transportweg in Richtung Lumen freigesetzt.
Article
Time-resolved absorption spectroscopy is a powerful tool to unravel biological functions and has been a key technology for elucidating the working of electron transfer chains in photosynthesis or photorepair of UV-damaged DNA. Both of these areas have seen important contributions from laboratories all over the world, not the least of them stemming from the ingenious technical advances described by Klaus Brettel, first at the Technical University of Berlin (Germany), and later at the Atomic Energy Agency in Saclay (France). Now, after more than forty years of tireless scientific activity, Klaus is approaching retirement and this collection gathers together tributes in the form of scientific contributions from colleagues along the way, covering a spectrum of topics as diverse as photosynthesis, light-induced DNA repair, electron and proton transfer in light signalling, flavin based photo-enzymology, fluorescent marker photophysics, synthetic models and modelisation, delicate sample transient absorption spectroscopy. In an era where science is increasingly changing context from "fundamental" to "applied", Klaus' curiosity and tenacity worked hand in hand in a most effective manner to further both technical possibilities and basic understanding.
Article
Thirty‐one and eleven sequences for the photosystem II reaction centre proteins D1 and D2 respectively, were compared to identify conserved single amino acid residues and regions in the sequences. Both proteins are highly conserved. One important difference is that the lumenal parts of the D1 protein are more conserved than the corresponding parts in the D2 protein. The three‐dimensional structures around the electron donors tyrosineZ and tyrosineD on the oxidizing side of photosystem II have been predicted by computer modelling using the photosynthetic reaction centre from purple bacteria as a framework. In the model the tyrosines occupy two cavities close to the lumenal surface of the membrane. They are symmetrically arranged around the primary donor P680 and the distances between the centre of the tyrosines and the closest Mg ion in P680 are around 14 A. Both tyrosineZ and tyrosineD are suggested to form a hydrogen bond with histidine 190 from the loop connecting helices C and D in the D1 and D2 proteins, respectively. The Mn cluster in the oxygen evolving complex has been localized by using known and estimated distances from the tyrosine radicals. It is suggested that a binding region for the Mn cluster is constituted by the lumenal ends of helices A and B and the loop connecting them in the D1 protein. This part of the D1 protein contains a large number of strictly conserved carboxylic acid residues and histidines which could participate in the Mn binding. There is little probability that the Mn cluster binds on the lumenal surface of the D2 protein.
Chapter
The Mn and Ca ions in Photosystem II are believed to be located at the interface between the intrinsic and extrinsic polypeptides of the PSII core. The recent identification of YZ [1,2] and YD [3,4] as Tyr-161 and Tyr-160 of the D1 and D2 polypeptides, respectively, suggests that many of the ligands to Mn and Ca are likely to be contributed by the D1 and D2 polypeptides. From considerations of Mn coordination chemistry[5,6], and from recent EPR [7] and electron spin echo [8] measurements with 15N, the Mn ions are believed to be coordinated primarily by carboxyl residues. Carboxyl residues are also likely ligands for Ca. There are 27 carboxyl residues in the lumenally-exposed regions of the mature forms of the D1 and D2 polypeptides. Several models for Mn and Ca ligation have been advanced incorporating a number of these residues [9,10]. However, there appears to be no compelling reason for selecting any particular residue over any other; unlike the previous situation with YZ and YD, there is no apparent C2 symmetry between the D1 and D2 polypeptides in this region.
Article
The oxygen flash yield and the kinetics of Chl a+II (P-680+) reduction have been measured under repetitive excitation as a function of pH between pH 4.0 and pH 9.0 in oxygen-evolving PS II particles from Synechococcus sp. (i) The optimum of oxygen yield is observed between pH 6.5 and pH 7.5. The inhibition in the acidic pH region is reversible and can be described by a monoprotic binding site with a pK value of about 4.5. In the alkaline pH region the inhibition is half maximal at pH 8.3 and might be described by the titration of three binding sites or more. The loss of oxygen evolution at pH 9.0 is caused by reversible inhibition and irreversible inactivation. (ii) Between pH 4.0 and pH 7.5 the fraction of Chl a+II decaying in the nanosecond time range and the oxygen yield follow the same pH dependence. (iii) Both in Photosystem II centers reversibly inhibited at low pH and in Photosystem II centers inactivated at high pH, Chl a+II is reduced by a donor Z, different from the normal immediate donor D1 or a modified state of D1, and, in part, by back reaction. (iv) Below pH 5.0, the decay in the nanosecond range can be explained by the existence of two phases with and (ratio of amplitudes, 1.3:1). A reduction phase with that is the major phase around the pH optimum is not observed below pH 5.0.
Article
In Tris-washed Photosystem-II particles we are able to induce an EPR signal in the dark by addition of an iridium salt (K2IrCl6). This signal is attributed to signal IIs (slow) (D+) and the redox titration gives an Em value of 760 mV for the couple . On the basis of our previous studies on the equilibrium between D+Z and DZ+ (K = 104) (Boussac, A. and Etienne, A.L. (1982) Biochem. Biophys. Res. Commun. 109, 1200–1205), we therefore attribute a value of 1 V for the Em of the couple. A second effect of K2IrCl6 is to modify the spectral characteristics of signal II. We conclude that K2IrCl6 is able to change the environment of the species from which signal IIs and signal IIf originate.
Article
In this report, we characterize the relationship between species “Z” (giving rise to EPR Signal II fast) and “D” (EPR Signal II slow) in triswashed chloroplasts.At pH 8.5 an externally added donor phenylenediamine competes with D for Z+ reduction after its oxidation by a flash. The reduction of Z+ by D occurs within some milliseconds. In a subsequent dark period, D+ is reduced by PD, the reaction rate being independant of phenylenediamine concentration. These results are consistent with the hypothesis of an equilibrium between Z+D and ZD+, the reduction of D by phenylenediamine occuring via Z. At lower pH's, the connection between Z and D is looser: a high concentration of phenylenediamine which reduces rapidly Z+, is very slow in reducing D+ and the subsequent photooxidation of D is less efficient.
Article
Photosynthetic water oxidation catalyzed by Photosystem II takes place at a site that comprises a redox-active tyrosine, YZ, a tetramanganese cluster, and, in addition to its redox components, two inorganic cofactors, calcium and chloride. Recent work suggests that YZ and the metal site are intimately linked in the oxidation and deprotonation reactions of substrate water. The metal cluster stores oxidizing equivalents and provides binding sites for the substrate from which YZ• is proposed to abstract hydrogen atoms during the catalytic cycle of photosystem II. Intrinsic to this hydrogen-abstraction mechanism for water oxidation is an intimate structural and functional relationship between the metal site and YZ, which predicts that the local YZ environment will be sensitive to the composition and integrity of the metal cluster. To test this postulate, we have examined the YZ site and its status with respect to solvent exposure under varying degrees of disassembly of the oxygen-evolving complex. 1H\2H-isotope exchange was carried out for various times in samples devoid of Mn, Ca2+, and Cl-, and in samples depleted exclusively of Ca2+. The YZ• and S2YZ• species were cryotrapped to high yield in these two preparations, respectively, and the radical site was characterized by using electron spin−echo envelope modulation spectroscopy. The isotope exchange at the YZ site was completed with an upper limit on the minutes time scale in both the (Mn)4-depleted and the Ca-depleted samples. The number of isotope-exchangeable protons in the site and their distances to YZ• were found to be different in the two systems, indicating that YZ is shielded from the solvent in the Ca-depleted system and, upon removal of the (Mn)4 cluster, becomes accessible to bulk water. The results from an electron spin−echo analysis of S2YZ•, in the weak-coupling limit, suggest that YZ• in samples that retain the (Mn)4 cluster, but lack Ca2+, is involved in a bifurcated hydrogen bond. The data for both classes of samples are consistent with a hydrogen-abstraction function for YZ in water oxidation and provide insight into the light-driven assembly of the (Mn)4 cluster.