ISSN 0006-2979, Biochemistry (Moscow), 2023, Vol. 88, No. 10, pp. 1580-1595 © Pleiades Publishing, Ltd., 2023.
Published in Russian in Biokhimiya, 2023, Vol. 88, No. 10, pp. 1908-1925.
1580
Femtosecond Dynamics of Excited States
of Chlorophyll Tetramer in Water-Soluble
Chlorophyll-Binding Protein BoWSCP
Dmitry A. Cherepanov
1,2,a
*, Konstantin V. Neverov
3,4
, Yuriy N. Obukhov
3
,
Yulia V. Maleeva
4
, Feodor E. Gostev
1
, Ivan V. Shelaev
1,2
, Arseny V. Aybush
1
,
Michail S. Kritsky
3
, and Victor A. Nadtochenko
1,5,b
*
1
Semenov Federal Research Center for Chemical Physics, Russian Academy of Sciences, 119991 Moscow, Russia
2
Belozersky Research Institute of Physical and Chemical Biology, Lomonosov Moscow State University,
119992 Moscow, Russia
3
Bach Institute of Biochemistry, Federal Research Center
“Fundamentals of Biotechnology” of the Russian Academy of Sciences, 119071 Moscow, Russia
4
Faculty of Biology, Lomonosov Moscow State University, 119991 Moscow, Russia
5
Faculty of Chemistry, Lomonosov Moscow State University, 119991 Moscow, Russia
a
e-mail: tscherepanov@gmail.com
b
e-mail: nadtochenko@gmail.com
Received June 22, 2023
Revised September 22, 2023
Accepted September 22, 2023
Abstract The paper reports on the absorption dynamics of chlorophylla in a symmetric tetrameric complex of the wa-
ter-soluble chlorophyll-binding protein BoWSCP. It was measured by a broadband femtosecond laser pump-probe spec-
troscopy within the range from 400 to 750nm and with a time resolution of 20fs-200ps. When BoWSCP was excited in
the region of the Soret band at a wavelength of 430nm, nonradiative intramolecular conversion S
3
→S
1
was observed with a
characteristic time of 83±9fs. When the complex was excited in the region of the Q
y
band at 670nm, relaxation transition
between two excitonic states of the chlorophyll dimer was observed in the range of 105±10fs. Absorption spectra of the ex-
cited singlet states S
1
and S
3
of chlorophylla were obtained. The delocalization of the excited state between exciton-coupled
Chl molecules in BoWSCP tetramer changed in time and depended on the excitation energy. When BoWSCP is excited in
the Soret band region, an ultrafast photochemical reaction is observed. This could result from the reduction of tryptophan
in the vicinity of chlorophyll.
DOI: 10.1134/S0006297923100139
Keywords: chlorophyll a, WSCP proteins, femtosecond laser spectroscopy, excited state spectrum, exciton dynamics, intra-
molecular conversion
Abbreviations: BoWSCP, water-soluble chlorophyll-binding
protein (WSCP) from Brassica oleracea var. botrytis; Chl,chlo-
rophyll.
* To whom correspondence should be addressed.
INTRODUCTION
Analysis of the photochemical reactions of energy
transfer and charge separation in photosynthetic pig-
ment–protein complexes is challenging due to their com-
plicated molecular organization and a large number
of exciton-coupled chlorophyll (Chl) molecules. Fur-
thermore, interpretating the absorption dynamics of
photosynthetic complexes registered by femtosecond
laser spectroscopy is a challenging methodological is-
sue [1, 2]. Water-soluble chlorophyll-binding pro-
teins (WSCP) from higher plants represent an ideal
model system to study the mechanism of photochem-
ical reactions [3-6]. These proteins are water-soluble,
photostable and thermally stable, and can be effective-
ly modified and produced through genetic and protein
engineering methods. In a plant cell these proteins are
FEMTOSECOND DYNAMICS OF CHLOROPHYLL TETRAMER 1581
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
located outside the thylakoid membranes of chloroplasts
and do not contribute to photosynthesis. Their physio-
logical functions are thought to be related to the activity
of the cells anti-stress systems [4, 7, 8].
Compared to the photosynthetic pigment–protein
complexes, WSCP protein structure is relatively simple.
WSCP holoforms are water-soluble homotetramers with
a molecular weight of 69-80 kDa. Each complex contains
up to four noncovalently bound Chl molecules. Depend-
ing on homology of the primary apoprotein structure,
WSCP proteins are subdivided into classesI and II, and
the latter has subclassesIIa and IIb by different affinity
of apoproteins for Chla and b [6,9]. It was shown that
upon photoexcitation of Chl in WSCP, energy migration
occurs both within the dimer and between dimer pairs
[10, 11]. This makes classII WSCP a promising model
for studying the mechanisms of pigment–pigment and
pigment–protein interactions, which will enable re-
search into the mechanisms of excitation energy transfer
between pigments [12-14] and interaction of the excited
pigments with surrounding proteins [15-17].
In this paper, a broadband femtosecond laser pump-
probe spectroscopy was used for studying photochemi-
cal processes in a symmetric tetrameric complex of the
water-soluble chlorophyll-binding protein BoWSCP
(subclassIIa) from Brassica oleracea var. botrytis, which
mainly binds Chla, i.e., its pigment composition can be
considered as almost homogeneous [5]. The tetrameric
Chl complex in BoWSCP is in the form of two dimers,
and the angle between porphyrine nuclei of pigments
in the dimer is about 30° [18]. Taken together, the two
dimers in a holoprotein macromolecule form a non-co-
valent tetrameric structure, where Chl dimer pairs are
connected by phytol tails [19]. The exciton interac-
tion energy of monomers in the dimer is approximate-
ly100 cm
–1
, while the interaction between dimers is an
order of magnitude weaker [13, 20]. It has recently been
shown that Chl forming part of this structure manifests
photocatalytic activity in redox processes [21].
This work aims to quantitatively analyse the pro-
cesses of nonradiative intramolecular conversion from
the third excited state to the lowest one, and electron-
ic transitions between exciton substates of the Q
y
band
of the Chl a tetramer when BoWSCP is excited in the
region of the Soret band and the Q
y
band, respective-
ly. Spectral characteristics are considered, which allow
the determination of the degree of exciton delocalization
between Chl molecules in the BoWSCP tetrameric com-
plex and the detection of electron transfer reactions in-
volving excited Chl molecules.
MATERIALS AND METHODS
BoWSCP apoprotein preparations were obtained
by expression of genes encoding this protein incorpo-
rated in a plasmid, by which cells of Escherichia coli
BL21(DE3) producer strain were transformed, in ac-
cordance with the protocol [5, 6, 22]. Due to six histi-
dine residues (6xHis) on N-terminal domains of apo-
proteins, their preparations were purified using affinity
chromatography on a Ni-agarose column with a subse-
quent preparation dialysis. Preparation purity was test-
ed using denaturing polyacrylamide gel electrophoresis
[5, 6]. Self-assembly of BoWSCP holoforms was per-
formed in vitro by joint incubation of apoproteins and
isolated thylakoid membranes of spinach [4-6, 8, 22].
Taking into account high thermal stability of WSCP
protein tetramers, the obtained holoform preparations
were heated (95°C, 10 min) to remove unnecessary pro-
teins. The final purification of BoWSCP was then car-
ried out on the Ni-agarose column [1,23]. The purified
preparation was dialysed against 10 mM Tris-HCl buf-
fer, pH 8.0. Quality of BoWSCP tetramer preparations
was tested using native polyacrylamide gel electropho-
resis, absorption spectroscopy, fluorescence spectros-
copy and circular dichroism spectroscopy. Denaturating
and native electrophoresis of proteins was carried out in
Mini PROTEAN Tetra System (Bio-Rad, USA). Pro-
tein content in samples was measured using the Brad-
ford assay [24]. The absorption spectra of the samples
were registered using Cary-50 (Agilent, USA) and
Shimadzu-1601PC (Shimadzu, Japan) spectrophotom-
eters. Circular dichroism (CD) spectra were registered
using Chirascan circular dichroism instrument (Applied
Photophysics, USA). Signal amplitude (Δε) was con-
verted into units of ellipticity(θ) for circular dichroism
spectra in accordance with [23].
Preparations for femtosecond spectroscopy were
concentrated by centrifugation in Amicon Ultra-15 cen-
trifugal devices with 30 or 50kDa filters. Concentration
was achieved by centrifugation of BoWSCP prepara-
tions (several times) in 10 mM Tris-HCl buffer, pH 8.0,
with addition of up to 150 mM NaCl or 10% glycerine
to prevent protein aggregation in the highly concentrat-
ed solution. The concentrate was diluted with the same
buffer with NaCl or glycerine to obtain 1ml of solution
with OD
673
= 5.
Ultrafast absorption changes ΔA(λ
i
,t
m
) were record-
ed by the femtosecond pump-probe laser spectroscopy
in the spectral range of 400 λ 780 nm and time de-
lays from 20 fs to 200 ps. The experimental setup and
measurement methodology were described earlier [25].
The excitation of the BowSCP was achieved by fem-
tosecond pulses with a maximum at 430 nm (duration
40 fs, energy 20 nJ) and 670 nm (duration 16 fs, energy
15 nJ). Absorption changes were enquired by the broad-
band pulses at the polarization angles of 0, 54.7, and
90 degrees relative to the exciting pulse polarization.
The difference absorption spectra in the range of 400-
780 nm were registered using the CCD-camera (Roper
Scientific SPEC-10, USA), connected to the polychro-
CHEREPANOV et al.1582
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
mator (Acton SP-300, USA). The experiments were car-
ried out at a temperature of 6°C in a circulating opti-
cal cell 0.5 mm thick, the diameter of optical windows
0.2 mm, the optical density of the sample 10 cm
−1
.
The rate of sample circulation (9 ml/min) was high
enough to avoid multiple excitations of the same sample
volume. The spectra were corrected by the group delay
dispersion, as previously described [26, 27]. Deconvo-
lution of the instrumental function accounting for the
coherent artifact in the region t = 0 was carried out ac-
cording to the method proposed in the works [26, 28].
For this purpose, the absorption changes caused by
nonlinear interactions of overlapping pump and probe
pulses (coherent spike) were approximated by the basic
gaussian function(1):
G(t) = 1/
2πd
2
exp[−½(t/d)
2
], (1)
and its derivatives in time G′ and G′′, where the width
of the pump pulse d was determined by the absorption
changes of the pure solvent [29]. The signal related to
changes in the electronic states of pigments was approx-
imated by convolution of the basic Gaussian function
G(t) and the step function of Hevisaid H(t) (2):
χ(t) = G(t)*H(t) = ½Erfc[−t/
2πd
2
], (2)
as well as convolutions of G(t) and exponential func-
tions(3):
ε(t,τ
k
) = G(t)*exp(−t/τ
k
) =
= ½exp[½(d/τ
k
)
2
t/τ
k
]Erfc[(d/τ
k
t/d)/
2].
(3)
The amplitudes A
k
i
) were determined by linear
regression and the characteristic times τ
k
by nonlinear
minimization.
The mathematical analysis of the absorption dynam-
ics included approximation of spectral-temporal ma-
trices ΔA(λ
l
,t
m
) by a linear combination of n discrete
exponential functions(4):
Q[τ](λ,t) =
Σ
D
k
(λ)exp(−t/τ
k
)+D
n+1
(λ), (4)
under the assumption that the characteristic times
τ={τ
k
} do not depend on the probe wavelength λ and,
therefore, can be considered as “global” parameters de-
termined by nonlinear minimization of the normalized
discrepancy (5):
R[τ] = (LMn)
−1
ΣΣ
A(λ
l
,t
m
)−
Q[τ](λ
l
,t
m
)]
2
.
(5)
The wavelength-dependent amplitudes D
k
(λ), which
were calculated by the linear regression method for each
exponential functionk, represent decay associated spec-
n
k = 1
L
l = 1
M
m = 1
tra(DAS). When the optimal values {τ
k
} of characteristic
times are found, the standard errors of their estimation
can be evaluated using the Jacobi matrixJ(6):
J
jk
=
δQ
j
[τ]
δτ
k
. (6)
Here the index j runs over L × M experimental
points of the spectral-temporal matrix ΔA(λ
l
,t
m
) [30].
The numerical assessment of Jacobian can be obtained
by decomposing the expression Q
j
[τ] in a Taylor series of
the first order (7):
δQ
j
[τ]
δτ
k
Q
j
[τ τe
k
]−Q
j
[τ −Δτe
k
]
τ
k
. (7)
Here τ is a set of optimal values of characteristic
times{τ
k
}, e
k
is the k
th
single vector, and Δτ is a small in-
crease of τ (the value of 0.015τ was used for a numerical
assessment). Standard errors σ
k
of characteristic times
estimation can be obtained from the diagonal elements
of the covariance matrix C=R[τ]×(J
T
J)
−1
(8):
σ
k
=
C
kk
. (8)
The values of σ
k
are standard deviations character-
izing the uncertainty of{τ
k
} assessments. According to
the Student’s t-distribution, 95% reliability is obtained
with the doubling of standard deviations, so the trust in-
tervals are τ
k
±2σ
k
, respectively.
RESULTS
Using the method proposed by Hughes etal. [20],
the pigment composition of the obtained BoWSCP ho-
loform preparations was determined by decomposing
its absorption spectrum in the Soret band region (360-
500 nm) into relative contributions of Chl a and Chl b
using the absorption spectra of individual pigments
in 90% acetone [31]. Figure 1a shows comparison of
BoWSCP absorption spectrum with the spectra of Chl a
and Chl b. In the absorption spectrum of free Chl a, the
Soret and Q
y
absorption bands with peaks at approxi-
mately 430 and 663 nm have a comparable amplitude;
in the spectrum of Chl b, these bands are located closer
to each other at 458 and 645 nm, respectively, and the
amplitude of the Soret band is increased, whereas that
of the Q
y
band is reduced in relation to the Chl a bands.
The presence of Chl b in the BoWSCP spectrum is de-
tected based on a small “arm” in the region of 470 nm,
while absorption in the interval of 380-430 nm is related
mainly to Chl a. Comparison of the spectra of BoWSCP
and free Chl a shows a significant effect of surround-
ing protein in the region of 360-450 nm. Firstly, Chl a
bands are shifted to the red by ~8 nm. Secondly, the in-
tensities of the two main transitions of Chl a in the Soret
region (the B
x
band with the peak at 430 nm corresponds
FEMTOSECOND DYNAMICS OF CHLOROPHYLL TETRAMER 1583
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
Fig. 1. Determination of the contributions of Chla and Chlb to the absorption spectrum of BoWSCP. a)Comparison of the absorption spectrum
of BoWSCP holoform in the buffer solution(1) with the spectra of Chla(2) and Chlb(3) in 90% acetone (based on the open spectral database
[31]). b)Modelling of BoWSCP spectrum(1) using Chla(2) and Chlb(3) spectra modified in accordance with the equation(9). In the re-
gion of 500-720nm, the amplitude of Chla spectrum is increased by 15%. Relative contributions of Chla and Chlb to the model spectrum(4)
are 89% and 11%, respectively.
to the S
0
→S
3
transition, and the B
y
band with the peak
at 380 nm corresponds to the S
0
→S
4
transition [32]) are
changed significantly in favour of the band with the peak
at ~380 nm. The electronic structure of the absorption
spectrum of Chl a in the Soret region cannot be de-
scribed properly by the Gouterman four-orbital mod-
el [33], because there are non-Gouterman states be-
tween the excited B
x
and B
y
states, mixed with vibronic
substates of the B
x
band. Consequently, energy and
dipole power of these states depend on dielectric sur-
rounding of the pigment [32, 34, 35].
As follows from Fig.1a, the relative probability of
S
0
→S
3
transition of Chl a in BoWSCP is reduced, while
the probability of S
0
→S
4
transition is increased compared
to the properties of free Chl a in solution. For this rea-
son, the BoWSCP absorption spectrum was modelled by
the empirical expression representing a sum of modified
absorption spectra of Chl a and Chl b, and taking into
account the redistribution of Chl a bands intensity in
the region of 360-500 nm:
A(λ) = αA
Chla
(λδ
а
)⋅[1 γtanh((λλ
а
/Δ)] +
+ βA
Chlb
(λδ
b
).
(9)
Here, α and β are the relative fractions of Chl a and
Chl b in the BoWSCP tetramer, A
Chla
and A
Chlb
are the
spectra of Chl a and Chl b in 90% acetone, δ
a
and δ
b
show how the spectra of Chl a and Chl b in BoWSCP
are shifted relative to their spectra in the solution.
The empirical function γ·tanh((λλ
a
)/Δ) describes
changes in the ratio of the B
x
and B
y
bands of Chl a
intensities in the spectral interval centered at the wave-
lengthλ
a
and the width Δ. Values of these parameters
determined by non-linear minimization amounted
to: α = 89.2%; β = 10.8%; δ
a
= 9.2 nm; δ
b
= 1.4 nm;
γ = 0.67; λ
a
= 413 nm; Δ = 50 nm; the model spectrum
CHEREPANOV et al.1584
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
Fig. 2. Location of chlorophyll molecules in the BoWSCP tetramer-
ic complex according to the X-ray crystallographic structure PDBID:
6s2z [18], where Chlb is replaced by Chla. Arrows show the directions
of vectors μ
k
(k = 1,…,4) characterizing the S
0
→S
1
transition dipole
moments of four Chl monomers, and dipole moments of four exciton
states E
n
of the tetrameric complex.
Fig. 3. The absorption spectrum (a), its second(b) and fourth(c) derivatives of BoWSCP holoform (dots). Solid lines represent a model spectrum,
and dashed lines show its main Gaussian components. Vertical dashed lines show the maxima of the fourth derivative of the absorption spectrum.
is shown in Fig. 1b by a thin solid line. Relative contri-
butions of Chl a and Chlb were equal to 89% and 11%,
respectively. The ratio between Chla/b pigments of 8 : 1
corresponds to the range of values 6-10 : 1 obtained in
various papers for BoWSCP native protein [3].
Absorption of BoWSCP in the Q
y
band region
(660-700 nm) is related primarily to Chl a, while the
Chl b contribution is negligible in this interval (Fig. 1b).
The BoWSCP spectral properties are adjusted by the
molecular structure of the pigment–protein complex:
four identical subunits form a tetramer in which three or-
thogonal axes of rotational symmetry C
2
(X)C
2
(Y)C
2
(Z)
define the coordinate system XYZ [18]. Four chloro-
phyll molecules in the hydrophobic locus form a dou-
ble dimer: interactions between dimers are much weak-
er than interactions between monomers within a dimer
(Fig.2).
The effect of exciton interaction between Chl mol-
ecules within the dimer predominates in the region of
660-700 nm, while the interaction between dimers is
much weaker. The shape of BoWSCP absorption spec-
trum in the Q
y
band region (Fig. 3a) indicates the presence
FEMTOSECOND DYNAMICS OF CHLOROPHYLL TETRAMER 1585
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
Fig. 4. Absorption changes of BoWSCP at the wavelengths of 450nm(a) and 670nm(b), induced by the pulse at 670nm with a duration of 16fs
(the black solid line). The coherent artifact was approximated by superposition of the Gaussian function G(t) (Eq.1) and its derivatives G′ and
G′′ (the dashed line), while the BoWSCP absorption changes were modelled by a combination of functions χ(t) and ε(t,τ
k
) (the dot-dash line),
with the parameter value d=24fs. The BoWSCP absorption response was obtained by subtraction of the coherent artifact (the red solid line).
of two exciton bands with peaks at 672 and 684 nm,
corresponding to the minima of the second derivative
(Fig. 3b) and maxima of the fourth derivative (Fig. 3c).
Using the method proposed for analyzing WSCP spec-
tra [20], the parameters of Chl a exciton interaction
were determined by decomposing the BoWSCP absorp-
tion spectrum in the Q
y
(0,0) band region into Gauss-
ian components. For an appropriate approximation
of the Q
y
(0,0) band with intramolecular vibrational
modes of the Chl tetrapyrrole macrocycle taken into
account, additional Gaussian components were used in
the high-energy region of the Q
y
(0,0) band [20, 36, 37].
The relative intensity of the main E
1
band peaking at
672 nm was 79%, while the contribution of the sec-
ond E
2
band peaking at 684 nm was about 21%. Spec-
tral characteristics of the E
3
state were not determined,
and the E
4
state has zero intensity in the approximation
ofpoint dipoles.
Broadband femtosecond laser pump-probe spec-
troscopy was used for measuring the absorption dynam-
ics of BoWSCP holoform, induced by excitation in the
region of wavelengths of 430 and 670 nm. Excitation in
these spectral regions is related mainly to Chl a (quan-
tum yield is >98%) and only marginally to Chl b (<2%).
CHEREPANOV et al.1586
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
Fig. 5. Transient absorption spectra of BoWSCP at time delays of 50fs(a) and 200ps(b) induced by excitation at 430nm(1) and 670nm(2).
The spectra are scaled in the Soret band region of 400-500nm at the delay of 200ps. The polarization angle is 54.7°. The inset shows the anisotro-
py coefficient in the Q
y
band region
Figure 4 shows the coherent artifact (coherent spike)
arising at short time delays due to nonlinear interactions
of overlapping pump and probe pulses. The contribution
of the coherent spike was divided from the absorption
changes of the pigments according to the procedure
described in “Materials and Methods”. The BoWSCP
absorption changes, starting from the time delay t = 0,
were assessed by deconvolution of the resulting signal
inaccordance with equations(2) and(3).
Figure 5a shows transient absorption spectra of
BoWSCP at a time delay of 50 fs, induced by the puls-
es at 430 nm (1,the solid line) and 670 nm (2,the dot-
dash line). The 430 nm excitation is related mainly to
the Chl transition S
0
→S
3
into the third excited singlet
state (the B
x
band), while the 670 nm excitation produc-
es the S
1
state (the Q
y
band). This assignment is con-
firmed by a negative anisotropy observed in the region
of the Q
y
band upon excitation at 430 nm, where the
anisotropy coefficient is close to the theoretical limit
of–0.1 (see the inset in Fig.5a).
The amplitude of the Q
y
band bleaching in the
BoWSCP transient absorption spectra in Fig. 5 is
3-5times higher than that of the Soret band, while the
amplitude of the Q
y
band in the BoWSCP linear absorp-
tion spectrum in Fig. 1a is comparable to that of the
Soret band. A transient absorption spectrum observed
FEMTOSECOND DYNAMICS OF CHLOROPHYLL TETRAMER 1587
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
Fig. 6. Decay associated difference spectra of BoWSCP in the range of 100 fs(a) and 10 ps(b) after excitation at 430nm (1,3) and 670nm (2,4).
Characteristic times: 1)83fs, 2)105fs, 3)19ps, 4)14ps. The marks on the graphs in the Q
y
band area show the confidence intervals calculated
using equation(8).
in the femtosecond measurements is the sum of three
contributions. First is the spectrum of bleaching, which
results from the disappearance of the ground state S
0
,
and its form corresponds to the inversed linear absorp-
tion spectrum. Second is the absorption spectrum of the
emerging excited state S
*
, and its contribution is always
positive. Third is the spectrum of the reverse S
*
→S
0
tran-
sition, stimulated by emission, and for the lowest excited
state its shape is close to the spontaneous fluorescence
spectrum; its contribution is always negative. The main
difference between the spectra (1) and (2) in Fig. 5a
is visible in the Q
y
band region, where the stimulated
emission of the S
1
→S
0
transition is observed, which does
not exist during the lifetime of the S
3
state produced by
the excitation at 430 nm. The less considerable differ-
ence between the spectra (1) and (2) in the region of
420-500 nm can be due to the stimulated emission of
the S
3
→S
0
transition.
Figure 5b shows transient absorption spectra at
a time delay of 200 ps using the same notations: 1)
excitation at 430 nm, 2) – 670 nm. Within this time
range the processes of S
3
→S
1
intramolecular conver-
sion and thermal equilibration between the four ex-
citon states E
n
of the tetrameric complex should be
completed. The observed final spectra are close in the
entire range, except for the Q
y
band, where the bleaching
amplitude after the excitation at 670 nm is much larg-
er than that of the bleaching induced by the excitation
at 430nm.
The most significant BoWSCP absorption changes
are related to the amplitude of the Q
y
band bleaching
after excitation at 430 nm. Its quantitative characteris-
tics were determined by decomposing the absorption
dynamics into exponential components (decay associ-
ated spectra, DAS, see “Materials and Methods”); de-
composition revealed spectral transitions in the range
of 100fs and 10ps (Fig.6).
Figure 6a shows the difference decay spectra for
the femtosecond range. After excitation at 430 nm the
main spectral changes occur with the characteristic
time of 83 ± 9 fs. They can be attributed to the nonra-
diative intramolecular S
3
→S
1
conversion, as a result of
which the stimulated emission disappears in the Soret
band region (450-480 nm) but appears in the region of
the Q
y
band (670-690 nm). When BoWSCP is excited
in the Q
y
band region at 670 nm, clearly distinguishable
small spectral changes occur with a characteristic time
of 105 ± 10 fs, and they can be attributed to the relax-
ation transition between the two excitonic substates E
1
and E
2
of the Q
y
band of Chl a dimer.
CHEREPANOV et al.1588
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
In the picosecond range, significant absorption
changes occur with a characteristic time of ~16 ps.
The DAS calculated for excitation at 430 nm (charac-
teristic time 19 ± 3 ps) and 670 nm (characteristic time
14 ± 2 ps) are close throughout the entire spectral range
(Fig. 6b) and are associated with insignificant absorp-
tion recovering in the region of the Q
y
band with a max-
imum around 675 nm. Possible interpretations of these
changes will be considered further.
To provide quantitative characteristics of photo-
chemical processes observed by means of femtosecond
measurements in the system of excitonically coupled
pigments, transient absorption (TA) was decomposed
into three components: ground state bleaching (GSB),
excited state absorption (ESA), and stimulated emis-
sion(SE). The probability of light absorption with the
quantum energy ħω
0
by a separate pigment molecule
is proportional to square of the dipole moment |μ|
2
for
its transition from the ground state to the excited state
(dipole strength of the transition). For Chl a monomer
in protein, the dipole moment of S
0
→S
1
transition is
equal to 5.5 D [38]. The pump pulse causes transition
of ΔN molecules in the sample from the ground state
to the excited state. The probe pulse is split into two
beams: one beam passes through a non-excited sam-
ple, and another beam passes through the excitation
region. The difference in beams intensity determines
the difference absorption. In the excitation region the
number of molecules in the ground state is decreased
by ΔN. Therefore, absorption at the frequency ω
0
in this
region is proportional to the transition dipole strength,
ΔA
GSB
|μ|
2
. At the same time, there is light emission of
the reverse S
1
→S
0
transition stimulated by the light wave,
its probability is also proportional to the dipole moment
strength, so the absorption at the frequency ω
0
is addi-
tionally decreased by the same value, ΔA
SE
|μ|
2
. For an
ideal two-level system, the ratio β = ΔA
SE
A
GSB
is equal
to one [39]; for Chl a in various solvents the estimate
was β≈0.9 [40,41].
For excitonically coupled pigments the situation is
way more complex. The dipole moment of excitonically
coupled pigments represents a linear sum of the dipole
moments of monomers [42]. In the BoWSCP crystal-
lographic structure [18], the dipole moments of mono-
mers μ
k
are almost parallel to the crystallographic axisX,
while projections of the moments to axes Y and Z are
much smaller in terms of absolute values [13]. In the first
approximation they may be assumed to be equal to zero.
Projections of the monomer dipole moment to crystallo-
graphic axes are shown schematically in Fig.2.
Quantitative description of light interaction with
Chl molecules in the BoWSCP tetramer raises a question
of basic nature: is the emerging excited state localized
within the dimer (the coupling energy of ~100 cm
–1
),
or is it delocalized across the entire tetrameric complex
(the coupling energy of ~10 cm
–1
)? The femtosecond
measurements enable determining the degree of excited
state delocalization, i.e., the number of Chl molecules
effectively involved in the exciton coupling.
For a symmetric dimer, the dipole moments of the
transitions to two possible excited states E
+
and E
are
equal [43] (10):
M
±
=
1
2
(μ
1
± μ
2
). (10)
It should be noted that the excitonic interaction of
monomers does not change the total dipole strength of
the complex: |M
+
|
2
+ |M
|
2
= μ
1
2
+ μ
2
2
, therefore, the in-
tegral amplitude of the linear absorption spectrum re-
mains constant. Due to the geometry of the BoWSCP
tetramer, the dipole strength of the transition to the E
+
state is equal to |M
+
|
2
μ
1
2
+ μ
2
2
= 2μ
1
2
, while the dipole
strength of the transition to the E
state is close to zero;
therefore, when BoWSCP is excited by a broadband
pulse centered at a wavelength of 670 nm (half-width of
50 nm, time duration of 16 fs), in the initial distribution
of arising exciton states, the E
+
state significantly pre-
dominates.
In this case, the total amplitude of bleaching ΔA
BL
in the initial transition spectrum near the exciton split-
ting band at frequency ω
0
is (11):
ΔA
BL
= ΔA
GSB
+ ΔA
SE
(1 + β)2μ
1
2
≈ 4μ
1
2
. (11)
In the rest of the spectral range, the bleaching of
the ground state is proportional to the dipole strength of
the free monomer (12):
ΔA
GSB
(ω) μ
1
2
. (12)
For a symmetric tetramer, the transition dipole
moments of four possible exciton states are found by a
unitary transformation [42], the moment with the larg-
est amplitude is equal to (13):
M
1
=
½
(μ
1
+ μ
2
μ
3
μ
4
). (13)
In this case, the total bleaching amplitude of the
transition spectrum near the exciton splitting band
is(14):
ΔA
BL
8μ
1
2
. (14)
In general, the number of excitonically coupled
molecules, n
excited
, between which the excited state is
delocalized, can be determined by the ratio of the total
bleaching amplitude of the transition spectrum at the
frequency of the exciton transition ω
0
and the amplitude
of the linear spectrum, which does not take into account
the selective nature of the complex excitation (15):
n
excited
= ½ΔA
BL
A
GSB
. (15)
FEMTOSECOND DYNAMICS OF CHLOROPHYLL TETRAMER 1589
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
Fig. 7. Decomposition of the BoWSCP transient absorption spectrum (1–Transient Absorption, TA) into components of ground state bleach-
ing (2–Ground State Bleaching, GSB), excited state absorption (3–Excited State Absorption, ESA) and bleaching in the stimulated emission
region (4–Bleach, BL). Transient absorption spectra were obtained at a delay of 50fs after excitation at 430nm(a) and 670nm(b). The bleach-
ing spectrum(2) was normalized based on the TA spectrum(1) in the region of 400-500nm, taking into account approximation of the excited
state spectrum by a cubic polynomial function(×). The model bleaching spectrum in the region of 650-700nm near the stimulated emission
(4) in the range ΔA
BL
=ΔA
GSB
A
SE
was represented by a sum ΔA
BL
=n
excited
·(ΔA
GSB
A
SE
) of the linear absorption spectrum and its mirror re-
flection relative to the maximum of the absorption band Q
y
peak and the Stokes shift Δλ. The model spectrum amplitude n
excited
was determined
by linear regression using polynomial approximation of the excited state spectrum(○). The thick gray line shows the result of such decomposition.
When BoWSCP is excited at 670 nm, the amplitude
of the ground state bleaching ΔA
GSB
can be determined
from the Soret band. In Fig. 7 dots show the transition
spectra of BoWSCP at a time delay of 50 fs after exci-
tation at wavelengths of 430nm(a) and 670 nm(b).
In the region of 400-500 nm, the transition spec-
tra (1) are close in shape to the absorption spectrum
of BoWSCP (2), with the exception of the region of
450-480 nm, where absorption of Chl b and stimu-
lated emission of Chl a from the S
3
state predominate.
Assuming that the absorption spectra of the excited
states S
1
and S
3
are smooth functions in the specified
wavelength range and can be correctly described by
third-degree polynomials P
3
(λ), the transition spectra
were approximated by the expression (16):
ΔA
TA
(λ) = ξ ⋅ΔA
GSB
(λ) + P
3
(λ), (16)
in which the scaling factor ξ and the coefficients of
the cubic polynomial P
3
were found by the standard
CHEREPANOV et al.1590
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
solution to the linear regression problem. The absorp-
tion spectra of excited states in the region where the
stimulated emission is absent are found as the differ-
ence ΔA
ESA
(λ) = ΔA
TA
(λ) − ξ⋅ΔA
GSB
(λ). Figure 7 shows
the obtained bleaching spectra of the ground state(2),
the polynomial functions P
3
(×), and the absorption
spectra of excited states(3).
The model bleaching spectrum near the Qy
band(4) was represented by the sum of the linear ab-
sorption spectrum taken with the opposite sign ΔA
GSB
(λ),
and its mirror reflection relative to the maximum of the
Qy band ΔA
SE
(λ). Since the position of the 0-0 transi-
tion remained unknown, the value of the Stokes shift Δλ
of the SE spectrum was found by nonlinear minimiza-
tion of the total model spectrum (17):
ΔA
TA
(λ) = n
excited
ξ⋅[ΔA
GSB
(λ) +
+ ΔA
GSB
(λ
0
+ Δλλ)] + P
3
(λ).
(17)
The effective number of excitonically coupled mol-
ecules n
excited
and the coefficients of the polynomial func-
tionP
3
were found by linear regression. Gray thick line
in Fig. 7 presents the result of this decomposition. When
excited at 430 nm, the observed Q
y
band bleaching
amplitude ΔA
BL
is only slightly larger than the ground
state bleaching amplitude ΔA
GSB
, while when excited at
670 nm, the ΔA
BL
amplitude is three times the ampli-
tudeΔA
GSB
.
Figure 8a shows time dependence of the bleach-
ing amplitude of the Soret band (1, 2) and the Q
y
band(3, 4), induced by BoWSCP excitation at 430 nm
(solid lines) and 670 nm (dot-dash lines). Absorption
changes in the time range of 100 fs with excitation at
430 nm (3) reflects stimulated emission which appears
as a result of electronic transition S
3
→S
1
, with a charac-
teristic time of 80 fs. As it was noted above, the bleach-
ing amplitude of the Q
y
band after excitation in the re-
gion of 430 nm remains much lower throughout the time
interval than the amplitude of this band after excitation
at 670 nm. The high bleaching amplitude of Q
y
band af-
ter excitation at 670 nm throughout the time interval is
another feature: its amplitude is three times higher than
that of the Soret band.
Figure 8b shows changes in the effective number
n
excited
of exciton-coupled Chl molecules in the course
of time, which is calculated according to the equa-
tion(15) for two BoWSCP excitation variants. Between
these molecules, the excited state of the tetramer is de-
localized. For short delays in the case of excitation at
670 nm, n
excited
is close to 2, i.e., BoWSCP acts as an ex-
citon-coupled dimer. In case of excitation at 430 nm at
short delays, n
excited
is close to 0.5. As it was noted above,
this means that the excited state S
3
prevails in these
conditions.
Finally, Fig. 8c shows changes in the anisotropy
coefficient, r = (ΔA
||
− ΔA
) / (ΔA
||
+ 2ΔA
), calculat-
ed in the Q
y
band region for two excitation variants.
The changes in the Q
y
band bleaching amplitude at
the timescale of ~100 fs are almost unrelated to the an-
isotropy changes.
DISCUSSION
WSCP proteins represent a uniquely organized
model system of exciton-coupled chlorophyll mole-
cules [11], which allows research into the mechanisms
Fig. 8. Changes in the main parameters characterizing the excited state
of BoWSCP tetrameric complex in the course of time. a) Bleaching
amplitude of the Soret band(1,2) and the Q
y
band(3,4) after exci-
tation at 430nm(1,3) and 670nm(2,4). b)The effective number of
excitonically coupled Chl molecules calculated by to the equation(15)
for excitation at 430nm(1) and 670nm(2). c)The anisotropy coeffi-
cient calculated in the Q
y
band region for excitation at 430nm(1) and
670nm(2).
FEMTOSECOND DYNAMICS OF CHLOROPHYLL TETRAMER 1591
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
of excitation energy transfer between pigments [12-14]
and the interaction of excited pigments with surround-
ing proteins [15-17]. Energy exchange between the
electron subsystem and nuclei vibration triggers the
evolution of electronic states [44, 45]. The efficiency
of energy exchange between exciton states and phonon
modes depends on spectral density of nuclei vibration-
al states. To determine this experimentally, spectral hole
burning [46-48] and time-resolved fluorescence spec-
troscopy [10] were used. Chl a vibron bands comprise
over 40 modes within the range of 1500 to 260 cm
−1
[49-51]. The dipole–dipole interaction of adjacent
Chl a molecules in the BoWSCP tetramer obtained by
joint decomposition of absorption and circular dichro-
ism spectra is ~100 cm
−1
[20], which allows considering
the delocalized excited states of pigments in terms of the
exciton theory [42]. The value of exciton interaction of
dimers of WSCP tetramer calculated using the X-ray
structure of the complex is ~10vcm
−1
[13], which allows
describing energy transfer between dimers using Förster
theory [52].
Measurement of absorption dynamics of WSCP
proteins of subclasses IIa (BoWSCP, from Brassica
oleracea) and IIb (LvWSCP, from Lepidium virginicum)
in the femtosecond range had been carried out in the Q
y
band region using pump-probe spectroscopy [14] and
2D spectroscopy [12, 13]. Theiss et al. measured absorp-
tion changes in the BoWSCP complex contained 20-30%
Chl b, and in that system, transfer of excitation from
Chl b to Chl a with a characteristic time of 400 fs was
registered [14]. That process was considered as an exci-
ton relaxation of heterodimer Chl a-Chl b, but later the
exciton relaxation of the heterodimer was shown to oc-
cur faster than in 100 fs, and processes with the time of
250-400 fs were ascribed to the energy transfer between
dimers [12]. The methods of 2D spectroscopy were used
for studying absorption dynamics of LvWSCP and
BoWSCP homotetramers containing 100% Chl a [13].
In both systems monoexponential absorption dynamics
with characteristic times of 70 and 100 fs, respectively,
were observed in the subpicosecond scale; these pro-
cesses were interpreted as the E
1
→E
2
relaxation to the
lowest exciton state [13]. Calculations with the modified
Redfield theory predict that equilibrium between exci-
ton states should be achieved within this time range [15].
The spectral transition with the time of 110 fs, observed
in case of BoWSCP excitation at 670 nm (Fig. 6a, spec-
trum2) matches the previously obtained data.
In this work the absorption dynamics of Chl a te-
trameric complex was first measured in the wide spec-
tral range of 400-750 nm with a time resolution of 20 fs.
This made is possible to register the intramolecular con-
version of Chl a from the excited state S
3
to S
1
, to ana-
lyze relaxation dynamics of exciton-coupled molecules
of Chl a, and to provide quantitative characteristics of
distribution between the E
1
and E
2
excitonic states relat-
ed to the Q
y
band of Chl a dimer. All that was possible
within a single experimental system.
Quantitative comparison of absorption dynamics in
the Soret and Q
y
band regions is required for properly
interpreting the data obtained with large photosynthetic
complexes. For instance, the bleaching amplitude of
Chl a in the Q
y
band region in primary transient ab-
sorption spectra of photosystemI from cyanobacterium
Thermosynechococcus elongatus in case of excitation in
the far-red region [53], and of Chl a and Chl f in similar
spectra of photosystem I from cyanobacterium Fisch-
erella thermalis PCC 7521 [54] is 3-6 times higher than
the bleaching amplitude of Soret bands. Meanwhile,
in similar spectra of photosystem I from cyanobacteri-
um Synechocystis sp. PCC 6803 the bleaching amplitude
of the Q
y
band is only slightly higher than that of the
Soret band [55]. Van Stokkum et al. assumed that the
high amplitude of the Q
y
band in the primary spectra
of photosystem I is an evidence for quenching of most
excited states by preoxidized chlorophyll of the special
pair P700 [56]. However, the light-harvesting complex-
es of photosystem I of cyanobacteria T. elongatus and
F. thermalis PCC 7521 contain exciton-coupled dimers
and trimers of Chl with maximum absorption in the far-
red region, as distinct from Synechocystissp. PCC 6803
[57-60]. The results obtained in this paper regarding the
impact of exciton coupling on the bleaching amplitude
of the Q
y
band (Fig.8) explain the observed effects by
structural features of the light-harvesting complexes.
Presented in Fig. 8, the dependencies of the bleach-
ing amplitude of the Soret band and the Q
y
band togeth-
er with the determined n
excited
coefficient representing the
number of exciton-coupled Chl molecules provide sev-
eral conclusions about the physical and chemical prop-
erties of BoWSCP excited states.
1. In case of BoWSCP excitation in the Q
y
band
region, the highest excitonic state E
1
with the peak at
672nm is predominantly formed. The thermal equilibri-
um between the E
1
and E
2
states (peak at about 684 nm)
is established at the characteristic time τ
1
= 105 ± 10 fs.
However, the contribution of the high-energy E
1
state
prevails over the low-energy E
2
state in the entire time
interval of hundreds of picoseconds (the relative contri-
bution is ≥80%). The population inversion can be due
to a considerable contribution of entropy to the differ-
ence in free energy of the E
1
and E
2
states. This impli-
cates a strong electronic interaction between BoWSCP
monomers, which takes place in case of considerable
overlapping of wave functions of excited states. In such
strongly coupled dimers, resonant charge-transfer states
arise, which determine a strong electron-phonon cou-
pling [61,62].
2. When BoWSCP is excited in the Soret band
region, the electronic state S
3
(also designated as B
x
)
is predominantly produced, which is transformed to
the S
1
state by a process of intramolecular conversion
CHEREPANOV et al.1592
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
with the characteristic time of 83 ± 9 fs. However, the
stimulated emission in the later state is characterized by
the amplitude which is almost halved compared to the
equilibrium S
1
state formed after excitation at 670 nm.
The decreased stimulated emission amplitude can re-
sult from an ultrafast photochemical reaction of Chl
oxidation and reduction of a tryptophan residue, “com-
peting” with the intramolecular conversion reaction.
In this case a reverse recombination reaction of Chl
+
Trp
ion-radical pair can take more than 200 ps. The pho-
toreduction of tryptophan Chl
*
Chl
+
Trp
was registered
for LvWSCP by spectral hole burning at the temperature
of 4 K in an hour timescale and with a low quantum
yield [47]. It can thus be assumed that spectral changes
with the characteristic time τ
2
= 16 ± 3 ps (Fig. 6b) can
result from acceleration of this photochemical reaction
at the temperature of 279 K.
3. An increased bleaching amplitude of the Q
y
band
in relation to the Soret band was observed in transient
absorption spectra of photosystemI of various cyanobac-
teria with specific excitation of long-wavelength chloro-
phyll forms [53, 54, 63]. Long-wavelength chlorophyll
forms in the light-harvesting antenna and the reaction
center of photosystem I arise due to formation of exciton-
ically coupled Chl dimers and trimers [60, 64, 65], which
obviously has the molecule structure similar to the archi-
tecture of chlorophyll molecules in the WSCP complex.
Contributions. D.A.Ch. analysis of spectral mea-
surement results; K.V.N., Yu.N.O., and Yu.V.M. pro-
duction, recovery, and biochemical characterization of
BoWSCP preparations; F.E.G., I.V.Sh., and A.V.A. de-
velopment of a mathematical apparatus, methods and
performing femtosecond measurements; M.S.K. and
V.A.N. concept development. All authors made a con-
siderable contribution to the manuscript.
Funding. This research was funded by the Russian
Science Foundation (grant no.21-74-20155).
Acknowledgments. The authors thank A. L. Dob-
ryakov for assistance in developing the coherent arti-
fact approximation method and setting the femtosecond
device. In this paper the equipment (the femtosecond
device) provided by the Center for Collective Use of
N. N. Semenov Federal Research Center for Chemi-
cal Physics, Russian Academy of Sciences “Analysis of
Chemical and Biological Systems and Natural Materi-
als: Mass Spectral Microscopy and Femtosecond Laser
Microscopy and Spectroscopy” (registration number:
506694). The expression vector pET-24b with the em-
bedded BoWSCP gene was kindly provided by Prof.
Dr. Harald Paulsen (Institute of Molecular Physiology,
Johannes Gutenberg University Mainz, Germany)
Ethics declarations. The authors declare no conflict
of interests in financial or any other sphere. This article
does not contain any studies with human participants
oranimals performed by any of the authors.
REFERENCES
1. Charlton, A., and Zachariou, M. (2007) Immobilized
metal ion affinity chromatography of histidine-tagged
fusion proteins, Methods Mol. Biol., 421, 137-149,
doi:10.1007/978-1-59745-582-4_10.
2. Cherepanov, D. A., Semenov, A. Y., Mamedov, M. D.,
Aybush, A. V., Gostev, F. E., Shelaev, I. V., Shuvalov,
V.A., and Nadtochenko, V.A. (2022) Current state of the
primary charge separation mechanism in photosystemI of
cyanobacteria, Biophys. Rev., 14, 805-820, doi: 10.1007/
s12551-022-00983-1.
3. Murata, T., Toda, F., Uchino, K., and Yakushiji, E. (1971)
Water-soluble chlorophyll protein of Brassica oleracea var.
botrys (cauliflower), Biochim. Biophys. Acta Bioenerg.,
245, 208-215, doi:10.1016/0005-2728(71)90023-5.
4. Satoh, H., Nakayama, K., and Okada, M. (1998) Molec-
ular cloning and functional expression of a water-soluble
chlorophyll protein, a putative carrier of chlorophyll mol-
ecules in cauliflower, J.Biol. Chem., 273, 30568-30575,
doi:10.1074/jbc.273.46.30568.
5. Takahashi, S., Yanai, H., Nakamaru, Y., Uchida, A.,
Nakayama, K., and Satoh, H. (2012) Molecular cloning,
characterization and analysis of the intracellular local-
ization of a water-soluble chl-binding protein from brus-
sels sprouts (Brassica oleracea var. gemmifera), Plant Cell
Physiol., 53, 879-891, doi:10.1093/pcp/pcs031.
6. Takahashi, S., Yanai, H., Oka-Takayama, Y., Zanma-
Sohtome, A., Fujiyama, K., Uchida, A., Nakayama, K.,
and Satoh, H. (2013) Molecular cloning, characterization
and analysis of the intracellular localization of a water-sol-
uble chlorophyll-binding protein (WSCP) from Virginia
pepperweed (Lepidium virginicum), a unique WSCP that
preferentially binds chlorophyll b in vitro, Planta, 238,
1065-1080, doi:10.1007/s00425-013-1952-7.
7. Satoh, H., Uchida, A., Nakayama, K., and Okada, M.
(2001) Water-soluble chlorophyll protein in Brassicace-
ae plants is a stress-induced chlorophyll-binding protein,
Plant Cell Physiol., 42, 906-911, doi:10.1093/pcp/pce117.
8. Horigome, D., Satoh, H., Itoh, N., Mitsunaga, K., Oon-
ishi, I., Nakagawa, A., and Uchida, A. (2007) Structural
mechanism and photoprotective function of water-soluble
chlorophyll-binding protein, J. Biol. Chem., 282, 6525-
6531, doi:10.1074/jbc.M609458200.
9. Maleeva, Y. V., Neverov, K. V., Obukhov, Y. N., and
Kritsky, M. S. (2019) Water soluble chlorophyll-bind-
ing proteins of plants: structure, properties and func-
tions, Mol. Biol. (Mosk), 53, 998-1011, doi: 10.1134/
S0026898419060120.
10. Schmitt, F. J., Trostmann, I., Theíss, C., Pieper, J.,
Renger, T., Fuesers, J., Hubrich, E. H., Paulsen, H.,
Eichler, H. J., and Renger, G. (2008) Excited state dy-
namics in recombinant water-soluble chlorophyll pro-
teins (WSCP) from cauliflower investigated by transient
fluorescence spectroscopy, J.Phys. Chem.B, 112, 13951-
13961, doi:10.1021/jp8024057.
FEMTOSECOND DYNAMICS OF CHLOROPHYLL TETRAMER 1593
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
11. Renger, G., Pieper, J., Theiss, C., Trostmann, I.,
Paulsen,H., Renger, T., Eichler, H.J., and Schmitt, F.J.
(2011) Water soluble chlorophyll binding protein of high-
er plants: Amost suitable model system for basic analy-
ses of pigment–pigment and pigment–protein interac-
tions in chlorophyll protein complexes, J.Plant Physiol.,
168, 1462-1472, doi:10.1016/j.jplph.2010.12.005.
12. Alster, J., Lokstein, H., Dostál, J., Uchida, A., and Zig-
mantas, D. (2014) 2D spectroscopy study of water-soluble
chlorophyll-binding protein from lepidium virginicum,
J.Phys. Chem.B, 118, 3524-3531, doi:10.1021/jp411174t.
13. Fresch, E., Meneghin, E., Agostini, A., Paulsen, H., Car-
bonera, D., and Collini, E. (2020) How the protein envi-
ronment can tune the energy, the coupling, and the ultra-
fast dynamics of interacting chlorophylls: the example of
the water-soluble chlorophyll protein, J.Phys. Chem. Lett.,
11, 1059-1067, doi:10.1021/acs.jpclett.9b03628.
14. Theiss, C., Trostmann, I., Andree, S., Schmitt, F. J.,
Renger, T., Eichler, H. J., Paulsen, H., and Renger, G.
(2007) Pigment–pigment and pigment–protein interac-
tions in recombinant water-soluble chlorophyll proteins
(WSCP) from cauliflower, J.Phys. Chem.B, 111, 13325-
13335, doi:10.1021/jp0723968.
15. Renger, T., Trostmann, I., Theiss, C., Madjet, M. E.,
Richter, M., Paulsen, H., Eichler, H.J., Knorr, A., and
Renger, G. (2007) Refinement of a structural model of a
pigment–protein complex by accurate optical line shape
theory and experiments, J. Phys. Chem. B, 111, 10487-
10501, doi:10.1021/jp0717241.
16. Friedl, C., Fedorov, D. G., and Renger, T. (2022) Towards
a quantitative description of excitonic couplings in photo-
synthetic pigment-protein complexes: Quantum chemistry
driven multiscale approaches, Phys. Chem. Chem. Phys.,
24, 5014-5038, doi:10.1039/d1cp03566e.
17. Lahav, Y., Noy, D., and Schapiro, I. (2021) Spectral tun-
ing of chlorophylls in proteins – electrostatics vs. ring
deformation, Phys. Chem. Chem. Phys., 23, 6544-6551,
doi:10.1039/d0cp06582j.
18. Agostini, A., Meneghin, E., Gewehr, L., Pedron, D.,
Palm, D.M., Carbonera, D., Paulsen, H., Jaenicke, E.,
and Collini, E. (2019) How water-mediated hydrogen
bonds affect chlorophyll a/b selectivity in water-solu-
ble chlorophyll protein, Sci. Rep., 9, 18255, doi:10.1038/
s41598-019-54520-4.
19. Bednarczyk, D., Dym, O., Prabahar, V., Peleg, Y., Pike,
D. H., and Noy, D. (2016) Fine tuning of chlorophyll
spectra by protein-induced ring deformation, Angew. Che-
mie Int. Ed., 55, 6901-6905, doi:10.1002/anie.201512001.
20. Hughes, J. L., Razeghifard, R., Logue, M., Oakley, A.,
Wydrzynski, T., and Krausz, E. (2006) Magneto-op-
tic spectroscopy of a protein tetramer binding two exci-
ton-coupled chlorophylls, J.Am. Chem. Soc., 128, 3649-
3658, doi:10.1021/ja056576b.
21. Obukhov, Y. N., Neverov, K. V., Maleeva, Y. V., and
Kritsky, M.S. (2023) Chlorophyll a dimers bound in the
water-soluble protein BoWSCP photosensitize the reduc-
tion of cytochromec, Dokl. Biochem. Biophys., 509, 60-64,
doi:10.1134/S1607672923700126.
22. Takahashi, S., Uchida, A., Nakayama, K., and Satoh,H.
(2014) Three-step photoconversion of only three sub-
units of the water-soluble chlorophyll-binding protein te-
tramer from Chenopodium album, ProteinJ., 33, 337-343,
doi:10.1007/s10930-014-9565-y.
23. Kelly, S. M., Jess, T. J., and Price, N. C. (2005) How to
study proteins by circular dichroism, Biochim. Biophys.
Acta Proteins Proteomics, 1751, 119-139, doi: 10.
1016/
j.bbapap.2005.06.005.
24. Bradford, M. M. (1976) Arapid and sensitive method for
the quantitation of microgram quantities of protein utiliz-
ing the principle of protein-dye binding, Anal. Biochem.,
72, 248-254, doi:10.1016/0003-2697(76)90527-3.
25. Cherepanov, D. A., Shelaev, I. V., Gostev, F. E., Mame-
dov, M.D., Petrova, A.A., Aybush, A.V., Shuvalov, V.A.,
Semenov, A.Y., and Nadtochenko, V.A. (2017) Mecha-
nism of adiabatic primary electron transfer in photosys-
temI: Femtosecond spectroscopy upon excitation of reac-
tion center in the far-red edge of the QYband, Biochim.
Biophys. Acta Bioenerg., 1858, 895-905, doi: 10.1016/
j.bbabio.2017.08.008.
26. Dobryakov, A. L., Pérez Lustres, J. L., Kovalenko, S. A.,
and Ernsting, N. P. (2008) Femtosecond transient ab-
sorption with chirped pump and supercontinuum probe:
Perturbative calculation of transient spectra with gener-
al lineshape functions, and simplifications, Chem. Phys.,
347, 127-138, doi:10.1016/j.chemphys.2007.11.003.
27. Golubeva, E. N., Zubanova, E. M., Melnikov, M. Y.,
Gostev, F. E., Shelaev, I. V., and Nadtochenko, V. A.
(2014) Femtosecond spectroscopy and TD-DFT calcu-
lations of CuCl
4
2−
excited states, Dalt. Trans., 43, 17820-
17827, doi:10.1039/C4DT01409J.
28. Dobryakov, A. L., Kovalenko, S. A., Weigel, A., Ṕrez-
Lustres, J.L., Lange, J., Müller, A., and Ernsting, N.P.
(2010) Femtosecond pump/supercontinuum-probe spec-
troscopy: Optimized setup and signal analysis for sin-
gle-shot spectral referencing, Rev. Sci. Instrum., 81,
113106, doi:10.1063/1.3492897.
29. Kovalenko, S. A., Dobryakov, A. L., Ruthmann, J., and
Ernsting, N. P. (1999) Femtosecond spectroscopy of
condensed phases with chirped supercontinuum prob-
ing, Phys. Rev. A At. Mol. Opt. Phys., 59, 2369-2384,
doi:10.1103/PhysRevA.59.2369.
30. Šimůnek, J. and Hopmans, J.W. (2002). 1.7 Parame-
ter Optimization and Nonlinear Fitting, in Methods of
Soil Analysis (eds Dane, J. H. and Clarke Topp, G.),
pp. 139-157, doi:10.2136/sssabookser5.4.c7.
31. Clementson, L. A., and Wojtasiewicz, B. (2019) Data-
set on the absorption characteristics of extracted phyto-
plankton pigments, Data Br., 24, 103875, doi: 10.1016/
j.dib.2019.103875.
32. Sirohiwal, A., Berraud-Pache, R., Neese, F., Izsák, R.,
and Pantazis, D. A. (2020) Accurate computation of the
absorption spectrum of chlorophyll a with pair natural
CHEREPANOV et al.1594
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
orbital coupled cluster methods, J. Phys. Chem.B, 124,
8761-8771, doi:10.1021/acs.jpcb.0c05761.
33. Gouterman, M. (1961) Spectra of porphyrins, J. Mol.
Spectrosc., 6, 138-163, doi:10.1016/0022-2852(61)90236-3.
34. Bricker, W. P., Shenai, P. M., Ghosh, A., Liu, Z., En-
riquez, M. G. M., Lambrev, P.H., Tan, H.S., Lo, C.S.,
Tretiak, S., Fernandez-Alberti, S., and Zhao, Y. (2015)
Non-radiative relaxation of photoexcited chlorophylls:
theoretical and experimental study, Sci. Rep., 5, 13625,
doi:10.1038/srep13625.
35. Götze, J. P., Anders, F., Petry, S., Witte, J.F., and Lok-
stein, H. (2022) Spectral characterization of the main
pigments in the plant photosynthetic apparatus by theory
and experiment, Chem. Phys., 559, 111517, doi: 10.1016/
j.chemphys.2022.111517.
36. Umetsu, M., Wang, Z. Y., Kobayashi, M., and Noza-
wa, T. (1999) Interaction of photosynthetic pigments
with various organic solvents. Magnetic circular di-
chroism approach and application to chlorosomes, Bio-
chim. Biophys. Acta Bioenerg., 1410, 19-31, doi: 10.1016/
S0005-2728(98)00170-4.
37. Shipman, L. L., Cotton, T. M., Norris, J. R., and Katz,
J. J. (1976) An analysis of the visible absorption spec-
trum of chlorophylla monomer, dimer,and oligomers in
solution, J.Am. Chem. Soc., 98, 8222-8230, doi:10.1021/
ja00441a056.
38. Cherepanov, D. A., Milanovsky, G. E., Aybush, A. V., and
Nadtochenko, V.A. (2023) Dipole moment of the S0-S1
chlorophyll a transition in solvents with varied refraction
index, Russ.J. Phys. Chem.B, 17, 584-593, doi:10.1134/
S1990793123030181.
39. Berera, R., van Grondelle, R., and Kennis, J. T. M. (2009)
Ultrafast transient absorption spectroscopy: principles
and application to photosynthetic systems, Photosynth.
Res., 101, 105-118, doi:10.1007/s11120-009-9454-y.
40. Cherepanov, D. A., Gostev, F. E., Shelaev, I. V., Aibush,
A. V., Mamedov, M.D., Shuvalov, V.A., Semenov, A.Y.,
and Nadtochenko, V.A. (2020) Visible and near infrared
absorption spectrum of the excited singlet state of chlo-
rophyll a, High Energy Chem., 54, 145-147, doi:10.1134/
S0018143920020058.
41. Cherepanov, D. A., Petrova, A. A., Mamedov, M. D.,
Vishnevskaya, A. I., Gostev, F. E., Shelaev, I. V., Aybush,
A. V., and Nadtochenko, V. A. (2022) Comparative ab-
sorption dynamics of the singlet excited states of chlo-
rophylls a and d, Biochem., 87, 1179-1186, doi: 10.1134/
S000629792210011X.
42. Davydov, A. S. (1962) The Theory Of Molecular Excitons,
McGraw-Hill, New York, NY.
43. Kasha, M., Rawls, H. R., and El-Bayoumi, M. A. (1965)
Theexciton model in molecular spectroscopy, Pure Appl.
Chem., 11, 371-392, doi:10.1351/pac196511030371.
44. Scholes, G. D., and Ghiggino, K. P. (1994) Rate ex-
pressions for excitation transfer I. Radiationless transi-
tion theory perspective, J. Chem. Phys., 101, 1251-1261,
doi:10.1063/1.467817.
45. Renger, T., and Marcus, R. A. (2002) On the relation of
protein dynamics and exciton relaxation in pigment-pro-
tein complexes: an estimation of the spectral density and
a theory for the calculation of optical spectra, J. Chem.
Phys., 116, 9997-10019, doi:10.1063/1.1470200.
46. Pieper, J., Rätsep, M., Trostmann, I., Schmitt, F. J.,
Theiss, C., Paulsen, H., Eichler, H.J., Freiberg, A., and
Renger, G. (2011) Excitonic energy level structure and pig-
ment-protein interactions in the recombinant water-solu-
ble chlorophyll protein. II.Spectral hole-burning experi-
ments, J. Phys. Chem. B, 115, 4053-4065, doi: 10.1021/
jp111457t.
47. Kell, A., Bednarczyk, D., Acharya, K., Chen, J., Noy, D.,
and Jankowiak, R. (2016) New insight into the water-sol-
uble chlorophyll-binding protein from Lepidium virgini-
cum, Photochem. Photobiol., 92, 428-435, doi: 10.1111/
php.12581.
48. Adolphs, J., Berrer, M., and Renger, T. (2016) Hole-burn-
ing spectroscopy on excitonically coupled pigments in
proteins: theory meets experiment, J. Am. Chem. Soc.,
138, 2993-3001, doi:10.1021/jacs.5b08246.
49. Gillie, J. K., Lyle, P. A., Small, G. J., and Golbeck,
J. H. (1989) Spectral hole burning of the primary elec-
tron donor state of PhotosystemI, Photosynth. Res., 22,
233-246, doi:10.1007/BF00048302.
50. Pieper, J., Voigt, J., and Small, G.J. (1999) Chlorophyll
a Franck–Condon factors and excitation energy transfer,
J.Phys. Chem.B, 103, 2321-2322, doi:10.1021/jp984460e.
51. Peternian, E. J. G., Pullerits, T., van Grondelle, R., and
van Amerongen, H. (1997) Electron-phonon coupling
and vibronic fine structure of light-harvesting com-
plexII of green plants: temperature dependent absorption
and high-resolution fluorescence spectroscopy, J. Phys.
Chem.B, 101, 4448-4457, doi:10.1021/jp962338e.
52. Förster, T. (1948) Intermolecular energy migration
and fluorescence [in German], Ann. Phys., 437, 55-75,
doi:10.1002/andp.19484370105.
53. Cherepanov, D. A., Shelaev, I. V., Gostev, F. E., Aybush,
A. V., Mamedov, M.D., Shuvalov, V.A., Semenov, A.Y.,
and Nadtochenko, V.A. (2020) Generation of ion-radical
chlorophyll states in the light-harvesting antenna and the
reaction center of cyanobacterial photosystem I, Photo-
synth. Res., 146, 55-73, doi:10.1007/s11120-020-00731-0.
54. Cherepanov, D. A., Shelaev, I. V., Gostev, F. E., Aybush,
A. V., Mamedov, M.D., Shen, G., Nadtochenko, V. A.,
Bryant, D. A., Semenov, A.Y., and Golbeck, J.H. (2020)
Evidence that chlorophyll f functions solely as an anten-
na pigment in far-red-light photosystemI from Fischerel-
la thermalis PCC7521, Biochim. Biophys. Acta Bioenerg.,
1861, 148184, doi:10.1016/j.bbabio.2020.148184.
55. Shelaev, I. V., Gostev, F. E., Mamedov, M. D., Sarkisov,
O. M., Nadtochenko, V.A., Shuvalov, V.A., and Semen-
ov, A. Y. (2010) Femtosecond primary charge separation
in Synechocystis sp. PCC 6803 photosystem I, Biochim.
Biophys. Acta Bioenerg., 1797, 1410-1420, doi: 10.1016/
j.bbabio.2010.02.026.
FEMTOSECOND DYNAMICS OF CHLOROPHYLL TETRAMER 1595
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
56. Van Stokkum, I. H. M., Müller, M. G., Weißenborn, J.,
Weigand, S., Snellenburg, J. J., and Holzwarth, A. R.
(2023) Energy transfer and trapping in photosystem I
with and without chlorophyll-f, IScience, 26, 107650,
doi:10.1016/J.ISCI.2023.107650.
57. Pålsson, L.-O. O., Flemming, C., Gobets, B., van Gron-
delle, R., Dekker, J.P., and Schlodder, E. (1998) Energy
transfer and charge separation in photosystemI: P700 oxi-
dation upon selective excitation of the long-wavelength an-
tenna chlorophylls of Synechococcus elongatus, Biophys.J.,
74, 2611-2622, doi:10.1016/S0006-3495(98)77967-6.
58. Gobets, B., Kennis, J. T. M., Ihalainen, J. A., Braz-
zoli, M., Croce, R., Stokkum, I. H. M. van, Bassi, R.,
Dekker, J.P., van Amerongen, H., Fleming, G.R., and
van Grondelle, R. (2001) Excitation energy transfer in di-
meric light harvesting complexI: acombined streak-cam-
era/fluorescence upconversion study, J. Phys. Chem. B,
105, 10132-10139, doi:10.1021/jp011901c.
59. Riley, K. J., Reinot, T., Jankowiak, R., Fromme, P., and
Zazubovich, V. (2007) Red antenna states of photosystemI
from cyanobacteria Synechocystis PCC 6803 and Ther-
mosynechococcus elongatus: single-complex spectroscopy
and spectral hole-burning study, J. Phys. Chem.B, 111,
286-292, doi:10.1021/jp062664m.
60. Gisriel, C., Shen, G., Kurashov, V., Ho, M.Y., Zhang, S.,
Williams, D., Golbeck, J.H., Fromme, P., and Bryant,
D.A. (2020) Thestructure of PhotosystemI acclimated
to far-red light illuminates an ecologically important accli-
mation process in photosynthesis, Sci. Adv., 6, eaay6415,
doi:10.1126/sciadv.aay6415.
61. Young, R. M., and Wasielewski, M. R. (2020) Mixed elec-
tronic states in molecular dimers: Connecting singlet fis-
sion, excimer formation, and symmetry-breaking charge
transfer, Acc. Chem. Res., 53, 1957-1968, doi: 10.1021/
acs.accounts.0c00397.
62. Casanova, D. (2018) Theoretical modeling of sin-
glet fission, Chem. Rev., 118, 7164-7207, doi: 10.1021/
acs.chemrev.7b00601.
63. Petrova, A. A., Casazza, A. P., Shelaev, I. V., Gostev,
F. E., Aybush, A. V., Nadtochenko, V. A., Semenov,
A. Y., Santabarbara, S., and Cherepanov, D. A. (2023)
Role of pheophytin a in the primary charge separation of
photosystem I from Acaryochloris marina: femtosecond
optical studies of excitation energy and electron trans-
fer reactions, Biochim. Biophys. Acta Bioenerg., 1864,
148984, doi:10.1016/j.bbabio.2023.148984.
64. Akhtar, P., and Lambrev, P. H. (2020) On the spectral
properties and excitation dynamics of long-wavelength
chlorophylls in higher-plant photosystem I, Biochim.
Biophys. Acta Bioenerg., 1861, 148274, doi: 10.1016/
J.BBABIO.2020.148274.
65. Karapetyan, N. V., Bolychevtseva, Y. V., Yurina, N. P.,
Terekhova, I. V., Shubin, V. V., and Brecht, M. (2014)
Long-wavelength chlorophylls in photosystemI of cyano-
bacteria: origin, localization, and functions, Biochemistry
(Moscow), 79, 213-220, doi:10.1134/S0006297914030067.