ArticlePDF Available

Unraveling the Electronic Structure of Individual Photosynthetic Pigment-Protein Complexes

Authors:

Abstract and Figures

Low-temperature single-molecule spectroscopic techniques were applied to a light-harvesting pigment-protein complex (LH2) from purple photosynthetic bacteria. The properties of the electronically excited states of the two circular assemblies (B800 and B850) of bacteriochlorophyll a (BChl a) pigment molecules in the individual complexes were revealed, without ensemble averaging. The results show that the excited states of the B800 ring of pigments are mainly localized on individual BChl a molecules. In contrast, the absorption of a photon by the B850 ring can be consistently described in terms of an excitation that is completely delocalized over the ring. This property may contribute to the high efficiency of energy transfer in these photosynthetic complexes.
Content may be subject to copyright.
Newcomb, and H. E. Dregne [Int. J. Remote Sens.15,
3547 (1994)]. SOI and SST data are produced and
archived by the National Oceanographic and Atmo-
spheric Administration/Climate Prediction Center
(http://www.cpc.ncep.noaa.gov/). A previous paper
(4) proposed the use of a satellite-derived potential
virus activity factor. We now use monthly NDVI data,
normalized to represent departures from the 1982–
95 mean, to better characterize rainfall anomalies
associated with RVF activity.
14. T. M. Logan, K. J. Linthicum, F. G. Davies, Y. S. Binepal,
C. R. Roberts, J. Med. Entomol.28, 293 (1991); T. M.
Logan, F. G. Davies, K. J. Linthicum, T. G. Ksiazek,
Trans. R. Soc. Trop. Med. Hyg.86, 202 (1992).
15. M. A. Cane, G. Eshel, R. W. Buckland, Nature 370, 204
(1994).
16. S. E. Nicholson, Int. J. Climatol. 17, 345 (1997);
iiii and J. Kim, ibid., p. 117.
17. There is a direct relation between rainfall and green
vegetation growth, between green vegetation growth
and the NDVI, and hence between rainfall and the
NDVI. This relation applies to areas receiving precip-
itation of ,800 mm/year. W. K. Lauenroth, in Per-
spectives in Grassland Ecology, N. French, Ed. (Springer-
Verlag, New York, 1979), pp. 3–24; H. N. Le Houerou
and C. H. Hoste, J. Range Manage.30, 181 (1977);
A. R. Malo and S. E. Nicholson, J. Arid Environ.19,1
(1990); S. E. Nicholson, M. L. Davenport, A. R. Malo,
Clim. Change 17, 209 (1990); C. J. Tucker and S. E.
Nicholson, Ambio, in press.
18. B. N. Holben, Y. J. Kaufman, J. D. Kendall, Int. J.
Remote Sens.11, 1511 (1990); E. F. Vermote and Y. J.
Kaufman, ibid. 16, 2317 (1995); S. O. Los, ibid. 14,
1907 (1994); N. Che and J. C. Price, Remote Sens.
Environ.41, 19 (1992).
19. Correlation coefficients were determined for a data
series calculated by the differences between adja-
cent values with SPSS Trends 6.1 software (SPSS,
Chicago, 1994). Nairobi NDVI anomalies were de-
rived from average monthly composite data within
8 by 8 grid cells, each with a spatial resolution of
8 km centered close to Nairobi, Kenya. Monthly
AVHRR data were derived from global area cover-
age data that are produced by the on-board pro-
cessing of large area coverage data (1.1 km by 1.1
km) and subsequently transmitted to receiving sta-
tions in Virginia or Alaska. Composite data were
formed by selecting the highest NDVI for each grid
cell location from daily data for that month to
minimize cloud and atmospheric contamination.
NDVI data were calculated and mapped to a Ham-
mer-Aitof projection. The highest value during a
monthly period was selected to represent the
monthly composite for each grid cell location.
20. AutoRegressive Integrated Moving Average (ARIMA)
analysis determined by SPSS, Trends 6.1 software.
21. T. M. Logan et al.,J. Am. Mosq. Control Assoc.6, 736
(1990).
5 February 1999; accepted 5 May 1999
Unraveling the Electronic Structure
of Individual Photosynthetic
Pigment-Protein Complexes
Antoine M. van Oijen,
1
* Martijn Ketelaars,
2
Ju¨rgen Ko¨hler,
1
Thijs J. Aartsma,
2
Jan Schmidt
1
Low-temperature single-molecule spectroscopic techniques were applied to a
light-harvesting pigment-protein complex (LH2) from purple photosynthetic
bacteria. The properties of the electronically excited states of the two circular
assemblies (B800 and B850) of bacteriochlorophyll a (BChl a) pigment mole-
cules in the individual complexes were revealed, without ensemble averaging.
The results show that the excited states of the B800 ring of pigments are mainly
localized on individual BChl a molecules. In contrast, the absorption of a photon
by the B850 ring can be consistently described in terms of an excitation that
is completely delocalized over the ring. This property may contribute to the high
efficiency of energy transfer in these photosynthetic complexes.
The primary process in bacterial photosynthesis
is the absorption of a photon by the light-
harvesting antenna system, followed by the rap-
id and efficient transfer to the reaction center
where the charge separation takes place. Typi-
cally, photosynthetic purple bacteria contain
two types of antenna complexes, light-harvest-
ing complexes 1 and 2 (LH1 and LH2, respec-
tively), both of which are integral membrane
proteins. The reaction center is presumed to be
surrounded by the LH1 complex, whereas the
LH2 complexes are arranged around the perim-
eter of the LH1 ring in a two-dimensional struc-
ture (1). The structure of the LH2 complex of
the purple bacterium Rhodopseudomonas aci-
dophila is known in great detail from x-ray
crystallography (2), which has shown that the
LH2 complex comprises 27 BChl a and (pre-
sumably) 18 carotenoid molecules nonco-
valently bound to the protein matrix. The BChl
a molecules are organized in two concentric
rings (Fig. 1). One ring, referred to as B800,
features a group of nine well-separated BChl a
molecules with an absorption band at ;800
nm. The other ring, referred to as B850, consists
of 18 closely interacting BChl a molecules with
an absorption band at ;860 nm. The entire
LH2 complex is cylindrically symmetric with a
ninefold symmetry axis. Upon excitation, ener-
gy transfer occurs from B800 to B850 mole-
cules on a picosecond time scale (3–5), whereas
among the B850 molecules, it is an order of
magnitude faster (6–8). The transfer of energy
from LH2 to LH1 and subsequently to the
reaction center occurs in vivo on a time scale of
5to25ps(9), very fast in comparison to the
decay of B850 in isolated LH2, which corre-
sponds to a lifetime of 1.1 ns.
Despite the fact that the LH2 complex has
been intensively investigated in recent years
with a wide variety of spectroscopic tools, in-
cluding the observation of the fluorescence dy-
namics of single LH2 complexes (10), no clear
picture of the electronic structure of its excited
states exists. Here, we present the results of a
study of isolated single LH2 complexes by
single-molecule fluorescence-excitation spec-
troscopy, a method successfully applied in re-
cent years to the detection of single guest mol-
ecules in crystalline and amorphous matrices
(11). This technique allows the observation of
optical spectra of individual complexes devoid
of the ensemble averaging over static intercom-
plex disorder, thus directly revealing the salient
properties of the electronic structure of the ex-
cited states.
The LH2 complexes of R. acidophila were
prepared as described elsewhere (3). Hydro-
lyzed poly(vinyl alcohol) (PVA) with a weight-
average molecular weight of 125,000 (obtained
from British Drug House) was purified over a
1
Centre for the Study of Excited States of Molecules,
2
Department of Biophysics, Huygens Laboratory, Lei-
den University, Post Office Box 9504, 2300 RA Leiden,
Netherlands.
*To whom correspondence should be addressed. E-
mail: antoine@molphys.leidenuniv.nl
Present address: Ludwig-Maximilians-Universita¨t
Mu¨nchen, Sektion Physik und Center for NanoScience,
Lehrstuhl fu¨r Photonik und Optoelektronik, Amalien-
strasse 54, 80799 Mu¨nchen, Germany.
zy
xx
Fig. 1. Geometrical ar-
rangement of the 27
BChl a molecules of
the LH2 complex of R.
acidophila obtained
by x-ray crystallogra-
phy. The B800 BChl a
molecules are depict-
ed in blue, and the
B850 pigments are
red. The phytol chains
of the BChl a mole-
cules are omitted for
clarity. The data have
been taken from the Protein Data Bank (identification code: 1kzu).
REPORTS
16 JULY 1999 VOL 285 SCIENCE www.sciencemag.org400
mixed resin in order to remove ionic impurities.
Thin polymer films, with a thickness of ,1
mm, were prepared by adding 1% w/w purified
PVA to a solution of 5 310
211
M LH2 in
buffer (0.1% lauryldimethylamine N-oxide, 10
mM tris, and 1 mM EDTA, with pH 8.0), which
was then spin coated on a LiF substrate (12).
The samples were mounted in a cryostat, cooled
to 1.2 K, and illuminated with a tunable con-
tinuous wave Ti-sapphire laser (spectral band-
width of 1 cm
21
). Microscopic images could be
obtained by wide-field illumination of the sam-
ple and imaging of the fluorescence at 890 nm
on a charge-coupled device camera. Fluores-
cence-excitation spectra were then acquired by
confocally exciting a spatially well-isolated
complex and detecting its fluorescence (also at
890 nm) with an avalanche photodiode. In both
cases, the detection bandwidth was 20 nm.
More details can be found in (13).
In Fig. 2, the fluorescence-excitation spectra
of several single LH2 complexes and an ensem-
ble of LH2 complexes are compared. The en-
semble spectrum features two broad structure-
less bands at ;800 and 860 nm, corresponding
to the absorptions of the B800 and B850 pig-
ments, respectively. When observing the single
complexes, the ensemble averaging in these
bands is removed, and remarkable spectral fea-
tures become visible. The striking differences
between the two absorption bands can be ratio-
nalized by considering the intermolecular inter-
action strength Jbetween neighboring BChl a
molecules in a ring and the spread in transition
energies D.Jis mainly determined by the inter-
molecular distance and the relative orientation
of the molecular dipole moments. Variations in
site energies Dcan often be attributed to struc-
tural variations in the environment of the BChl a
molecules, resulting in changes in the electro-
static interaction with the surrounding protein. If
the ratio J/Dis small, it is expected that the
excitations are mainly localized on individual
BChl a molecules. If the coupling strength J
between the BChl a molecules is much larger
than D, the description should be in terms of
delocalized excited-state wave functions with
relatively short energy relaxation times.
As can be seen in Fig. 2, the B800 band of
an individual LH2 complex consists of several
relatively narrow spectral lines. From the width
of the B800 ensemble line, a value of ;125
cm
21
for the diagonal disorder Dcan be ex-
tracted, and from the x-ray structure, it can be
calculated, with a point-dipole approximation,
that the interaction energy Jbetween neighbor-
ing pigments amounts to 224 cm
21
(14). The
ratio ?J/D?'0.2 is characteristic for electroni-
cally excited states, which are largely localized
on individual pigments. Therefore, the narrow
lines around 800 nm can be attributed to the
absorptions of individual BChl a molecules in
the B800 ring. This interpretation is corroborat-
ed by the strong dependence of the relative
intensities of these lines on the polarization of
the incident radiation, consistent with the dif-
ferent directions of the dipole moments of lo-
calized transitions of BChl a molecules in the
ring (15).
In the B850 band, the interaction strength
between the BChl a molecules is determined to
be ;300 cm
21
(14), that is, considerably larger
than the disorder (estimated to be ;125 cm
21
).
Therefore, we have to consider excitonic inter-
actions in order to understand the optical spec-
tra. As a starting point, we calculated the excit-
ed-state manifold of a cylindrically symmetric
B850 assembly with zero disorder. Of the two
nondegenerate (denoted as k50 and k59)
and eight pairwise degenerate (k561, k5
62, ..., k568) exciton states, only the
low-energy degenerate pair k561 will carry
appreciable oscillator strength (Fig. 3A, left).
Upon introducing diagonal disorder in the ring,
the pairwise degeneracies will be lifted, and the
oscillator strength is redistributed over adjacent
exciton states (16) (Fig. 3A, right). The transi-
tion dipole moments associated with the k5
61 transitions will have orthogonal polariza-
tions. This orthogonality is maintained when
disorder is introduced, assuming that the diag-
onal disorder is dominated by variations in
electrostatic interactions and possibly intermo-
lecular distances, rather than by changes in the
orientations of the BChl a molecules.
Fig. 2. Comparison of fluorescence-excitation
spectra for an ensemble of LH2 complexes (top
trace) and several individual LH2 complexes at 1.2
K. The vertical scale applies to the bottom spec-
trum; all other spectra are offset for clarity. cps,
counts per second.
k
kk
kk
AC
B
Fig. 3. (A) Schematic representation of the energy-level scheme of the lowest states in the excited-state
manifold of the B850 ring in LH2 of R. acidophila. Compared are the relative positions of the lowest
levels in the presence (left) and in the absence (right) of ninefold rotational symmetry. The gray circles
indicate the initial population of a given excited state, and the arrows indicate the relative orientation
of the transition dipole moments in the plane of the ring. (B) Fluorescence-excitation spectrum of the
long-wavelength region of an individual LH2 complex for mutually orthogonal polarized excitation as
schematically indicated in (A) by the colored arrows. (C) Fluorescence-excitation spectrum of the red
wing of the long-wavelength absorption in the B850 band. In the bottom panel, a stack of 200
consecutively recorded spectra (3 s per scan) is shown where the fluorescence intensity is given by the
color code (yellow corresponds to high intensities). The spectrum in the top panel corresponds to an
average of only those scans that are covered by the box. For this particular complex, the whole set of
lines in the B850 band is shifted toward higher energies in comparison to the complex shown in (B).
REPORTS
www.sciencemag.org SCIENCE VOL 285 16 JULY 1999 401
In all spectra of the single LH2 complexes
we observed, the B850 band consisted of two
broad absorption lines at ;860 nm, some-
times accompanied by a weaker third transi-
tion at the higher energy side. These obser-
vations can be explained in terms of the
exciton model. The two absorptions corre-
spond to the k561 transitions, with their
degeneracy lifted. By performing polariza-
tion-dependent experiments on these two
bands (17), the orthogonality of the associat-
ed transition dipole moments, predicted by
the exciton model, could be ascertained (Fig.
3B). This orthogonality was observed in all
individual LH2 complexes that we studied
and is a strong indication for a high degree of
delocalization of the excitation. The observed
homogeneous linewidth of the k561 tran-
sitions of ;50 cm
21
is consistent with an-
isotropy decay times of ;100 fs found in
pump-probe experiments (8). The extent of
delocalization will decrease, and the dynam-
ical properties will change at a higher tem-
perature, where mixing of the exciton states
by vibronic coupling will occur (18).
Another observation supporting the ex-
citonic level scheme is the detection of the
lowest exciton state k50. By repeatedly
scanning the excitation wavelength quickly
through the low-energy side of the k561
pair and following the spectral features
through time, the presence of the spectrally
rapidly diffusing lowest exciton state could
be made visible in a fraction (;25%) of the
studied complexes (Fig. 3C). The low in-
tensity of the k50 transition, which in
principle is dipole forbidden, and spectral
diffusion on a time scale faster than that of
the experiment explain the absence of this
lowest exciton transition in most of the
complexes. The linewidth of the k50 state
should be ;0.005 cm
21
, as determined by
the 1.1-ns fluorescence lifetime of the sys-
tem, but the observed value of ;5cm
21
is
mainly determined by residual spectral dif-
fusion and the bandwidth of the excitation
source.
The energy splitting dE
61
between the
k511 and k521 states was measured for
all complexes investigated (Fig. 4). As men-
tioned previously, the presence of the disor-
der Dforms a plausible cause for the lifting of
the degeneracy of these exciton states. How-
ever, simulations show that the observed av-
erage dE
61
of ;110 cm
21
cannot be ex-
plained by taking into account only random
disorder (as depicted in Fig. 4) for a simulat-
ed Dwith a full width at half-maximum
(FWHM) of 125 cm
21
(19). Even an unrea-
sonably large value of D'500 cm
21
for the
width of the distribution of site energies did
not result in an energy separation between the
k561 exciton states, as observed experi-
mentally. To exclude the possibility that
these abnormally high splittings are caused
by an anisotropic environment in the poly-
mer matrix, we repeated the experiment on
single LH2 complexes in a glycerol matrix,
which resulted in similar values of dE
61
.A
Jahn-Teller–like deformation in the excited
state can probably be ruled out in view of
the unrealistically high values needed for
the electronic-nuclear coupling strength.
Although random disorder can give rise to
large variations in the absorption wave-
lengths of the B850 bands (Fig. 2), the ob-
served energy separation of the k561 states
can only be explained in terms of largely
correlated disorder, such as a static symmet-
ric distortion of the protein complex in the
ground state. In the case of an elliptical de-
formation of the ring, it can be shown, on the
basis of symmetry arguments, that only the
k561 exciton states will be split. This is
consistent with the absence in the spectra of a
splitting and a polarization effect of the k5
62 states. The eigenfunctions of the k561
states belong to the long and short axes of the
ellipse and hence exhibit orthogonal polariza-
tion of their transition moments, as observed
experimentally. Our simulations show that
the observed splittings can be explained by
assuming an eccentricity «of the ring of 0.52,
corresponding to a ratio of the long and short
radius of 0.85, and a random disorder of 125
cm
21
(Fig. 4); «5(1–a
2
/b
2
)
1/2
, where aand
bare the length of the short and long axes,
respectively. An explanation for the sym-
metry lowering in the LH2 from ninefold in
the crystals used for resolving the x-ray
structure to the twofold symmetry observed
in our experiments may be found in the
extremely dense packing of LH2 in the x-ray
crystals, causing a stabilization of the struc-
ture. In our case of completely isolated com-
plexes, these stabilizing forces are absent,
and the complex deforms. What the symme-
try properties of the LH2 are in a natural
environment, surrounded by a limited num-
ber of LH2 complexes in the photosynthetic
membrane, is therefore an intriguing question
and deserves further study.
This work demonstrates that single-mole-
cule spectroscopy is a powerful tool to reveal in
detail the factors determining the electronic
structure of pigment-protein complexes and,
more generally, of molecular aggregates. Vari-
ous manifestations of disorder can be probed
directly, providing valuable information for the
theoretical modeling of energy-transfer pro-
cesses in these systems, a better understanding
of the structure of these biologically important
systems, and an understanding of how these
systems function.
References and Notes
1. M. Z. Papiz et al.,Trends Plant Sci. 1, 198 (1996).
2. G. McDermott et al.,Nature 374, 517 (1995).
3. J. T. M. Kennis, A. M. Streltsov, H. Permentier, T. J.
Aartsma, J. Amesz, J. Phys. Chem. B 101, 8369 (1997).
4. R. Monshouwer, I. Ortiz de Zarate, F. van Mourik, R.
van Grondelle, Chem. Phys. Lett. 246, 341 (1995).
5. H.-M. Wu et al.,J. Phys. Chem. 100, 12022 (1996).
6. M. Chachisvilis, O. Ku¨hn, T. Pullerits, V. Sundstro¨m, J.
Phys. Chem. B 101, 7275 (1997).
7. R. Jimenez, S. R. Dikshit, S. E. Bradforth, G. R. Fleming,
J. Phys. Chem. 100, 6825 (1996).
8. S. I. E. Vulto, J. T. M. Kennis, A. M. Streltsov, J. Amesz,
T. J. Aartsma, J. Phys. Chem. B 103, 878 (1999).
9. V. Nagarajan and W. W. Parson, Biochemistry 36,
2300 (1997).
10. M. A. Bopp, Y. Jia, L. Li, R. J. Cogdell, R. M. Hochstrasser,
Proc. Natl. Acad. Sci. U.S.A. 94, 10630 (1997).
11. Th. Basche´, W. E. Moerner, M. Orrit, U. Wild, Single
Molecule Optical Detection, Imaging and Spectrosco-
py (Verlag-Chemie, Munich, 1997).
12. LiF is a favorable substrate because the inversion
symmetry of the alkali halide crystal prevents first-
order Raman scattering.
13. A. M. van Oijen, M. Ketelaars, J. Ko¨hler, T. J. Aartsma,
J. Schmidt, J. Phys. Chem. B 102, 9363 (1999).
14. K. Sauer et al.,Photochem. Photobiol. 64, 564 (1996).
15. A. M. van Oijen, M. Ketelaars, J. Ko¨hler, T. J. Aartsma,
J. Schmidt, Chem. Phys., in press.
16. R. G. Alden et al.,J. Phys. Chem. B 101, 4667 (1997).
17. The polarization-dependent experiments were per-
formed by rotating the polarization of the incident laser
light in the plane of the sample. As a consequence of
the high-velocity spin coating, the LH2 complexes are
oriented predominantly flat on the substrate surface,
perpendicular to the propagation vector of the excita-
tion light. This was confirmed by similar experiments on
LH2 complexes in non–spin-coated films.
18. J. A. Leegwater, J. Phys. Chem. 100, 14403 (1996).
19. Monte Carlo simulations were performed on the basis
of the crystal structure of LH2 of R. acidophila. Only
nearest neighbor interactions were taken into ac-
count with the dipole-dipole approximation. A ran-
dom diagonal disorder (FWHM 5125 cm
21
) was
introduced for both the undistorted and distorted
rings, assuming that it is centered at the same tran-
sition energy for every B850 pigment. In the case of
the distorted ring, the individual pigments were po-
sitioned on an ellipse, whereas the long-wavelength
(Q
y
) transition dipoles retained the alignment as seen
in the crystal structure.
20. The authors thank J. P. Abrahams (Leiden University,
Leiden, Netherlands), J. Knoester (Groningen Univer-
sity, Groningen, Netherlands), and J. H. van der Waals
(Leiden University, Leiden, Netherlands) for many
helpful discussions. We also thank D. de Wit for the
preparation of the LH2 complexes and M. Hessel-
berth for assistance with the spin coating. This work
is supported by the Stichting voor Fundamenteel
Onderzoek der Materie (FOM) with financial aid from
the Nederlandse Organisatie voor Wetenschappelijk
Onderzoek (NWO). J.K. is a Heisenberg fellow of the
Deutsche Forschungsgemeinschaft.
8 April 1999; accepted 2 June 1999
Fig. 4. The distribution of the energy separations
dE
61
of the k561 transitions. The histogram
represents the experimental data (exp) for all
complexes studied. The values obtained by nu-
merical calculations (19), assuming only a disor-
der of 125 cm
–1
, are depicted by solid black
circles. After an additional elliptic deformation,
with an eccentricity of «50.52, is introduced in
the simulations (sim), one obtains the data rep-
resented by solid black squares.
REPORTS
16 JULY 1999 VOL 285 SCIENCE www.sciencemag.org402
... The morphology of the artificial nanomicelle system was first studied by transmission electron microscopy (TEM) and scanning transmission electron microscopy (STEM), which revealed that all ZnMSA1-4 formed ultrauniform spherical assemblies Nalg-1-4 ( Fig. 2a,b and Supplementary Figs. [1][2][3][4][5][6]. High-resolution STEM of Nalg-4 revealed a ring-like substructure with a diameter around 4.1 nm on its surface (Fig. 2c). ...
... Time-dependent GC experiments were performed to track the content of CO 2 , CO and CH 4 in the headspace for a Nalg-4 solution ([Zn] = 0.2 mM) with 2 μM of C2 and 20 mM of TEA·HCl (5% TEA) from an atmosphere (1,000 ppm CO 2 and 1,000 ppm CO in N 2 ) under visible light in water over 72 h (Supplementary Fig. 23). The concentration of CO reached 1,083 ppm in the first 5 h and then reduced at a slower rate compared with CO 2 over the first 12 h irradiation, which was accompanied with the production of CH 4 . This indicates that the CO is an intermediate product and could be further reduced to CH 4 in this nanomicelle system ( Supplementary Fig. 23). ...
... Ultrahigh-purity CO 2 (99.999%) was used as a carrier gas for CO and CH 4 detection, whereas ultrahigh-purity argon was utilized for H 2 detection. Initially, the GC system was calibrated for H 2 , CO and CH 4 . To confirm that the CO and CH 4 products were derived from CO 2 , an isotope 13 CO 2 (Sigma Aldrich) was used as the atmosphere gas for visible-light-irradiation experiments, and GC-mass spectroscopy was used for gas detection. ...
Article
Full-text available
In nature, photosynthetic organelles harness solar radiation to produce energy-rich compounds from water and atmospheric CO2 via exquisite supramolecular assemblies. Although artificial photocatalytic cycles have been shown to occur at higher intrinsic efficiencies, the low selectivity and stability in water for multi-electron CO2 reduction hamper their practical applications. The creation of water-compatible artificial photocatalytic systems mimicking the natural photosynthetic apparatus for selective and efficient solar fuel production represents a major challenge. Here we show a highly stable and efficient artificial spherical chromatophore nanomicelle system self-assembled from Zn porphyrin amphiphiles with a Co catalyst in water for CO2-to-methane conversion with a turnover number >6,600 and 89% selectivity over 30 days. The hierarchical self-assembly induced a spherical antenna effect that could facilitate the photocatalytic process with an initial 15% solar-to-fuel efficiency. Furthermore, it has a capability to efficiently reduce atmospheric CO2 into methane with high selectivity in water.
... Amongst various photoinduced processes, the light-harvesting (LH) properties of macroor supramolecular systems hold our attention in this work. Natural and synthetic systems may show LH properties [7][8][9][10][11][12][13][14][15][16][17][18][19][20][21][22]. Studies have been made on the link between the structure of the systems and the excitation energy transfer (EET) process [23][24][25][26][27][28][29]. ...
... Studies have been made on the link between the structure of the systems and the excitation energy transfer (EET) process [23][24][25][26][27][28][29]. Concerning natural systems, one can name the case of LH properties that arises from photosynthetic abilities of pigment-protein complexes [7][8][9][10][11]. ...
... The systems mentioned in the previous refs [7][8][9][10][11] are pigment-protein complexes of the purple bacteria. It is one of the first studied systems that show LH properties but it is not the only one. ...
Thesis
Full-text available
Phenylene-ethynylene dendrimers show astonishing properties and are systems to be likely used in opto-electronic devices such as light emitting diodes and conductive molecular wires. They are in the spotlight because they show excellent photostability and high excitation energy transfer efficiency. The excitation energy transfer in phenylene-ethynylene dendrimers is ultrafast and unidirectional. It occurs from the periphery of the molecular system to the core thanks to an excitation energy gradient that extends along the system.During this PhD, phenylene-ethynylene dendrimers have been studied through a pseudofragmentation scheme in which the phenylene-ethynylene dendrimers is decomposed in various subsystems (pseudofragments). The phenylene-ethynylene dendrimers behaves as if the pseudofragments (oligophenylene-ethynylene) were weakly interacting together.Two isomers (the single-trans isomer and the cumulenic isomer) of oligophenylene-ethynylene co-exist in their first adiabatic electronic excited states.Two diabatic excited states are then considered for each oligophenylene-ethynylene, the ones which are associated to the Lewis structures of the isomers. The potential energy surfaces of phenylene-ethynylene dendrimers and their conical intersections have been rationalised in terms of diabatic states localised on the pseudoframgments.I have used density-based descriptors that are built from the attachment and detachment densities involved in the electronic transitions. Such descriptors are used to characterise the electronic excited states that are involved in the pseudofragmentation scheme of phenylene-ethynylene dendrimers .This global strategy allowed us to suggest an alternative excitation energy transfer mechanism that involves both trans-bending and cumulenic-streching deformations on each of the pseudofragments of a phenylene-ethynylene dendrimers .
... 1−3 The strong electronic coupling (∼300 cm −1 ) between assembled chromophores in LH2 permits electronic coherence between chromophores and the formation of excitons composed of as little as one and as many as 18 molecular units (at low temperatures). 4 Subsequent energy transfer between these delocalized sites occurs either by incoherent hopping or by wave-like coherent transport, 5 enabling rapid and directional transfer of excitations to the reaction center, where charge separation takes place. ...
... This exciton "self-trapping" often occurs on the femtosecond time scale. 2,40 For example, the exciton size in B850 (a ring-like assembly of 18 strongly interacting bacteriochlorophyll chromophores found in photosynthetic purple bacteria) was studied using single-molecule spectroscopy at T = 1.2 K. 4 Under such conditions, the fluorescence excitation spectra of B850 indicate that the exciton delocalizes over the entire ring, which is consistent with strong electronic coupling (∼300 cm −1 ) and weak disorder (∼125 cm −1 ). At room temperature, the exciton is initially delocalized over 5− 13 molecules, while it rapidly (within 100 fs) decoheres into exciton size of 4 molecular units. ...
Article
Exciton size and dynamics were studied in assemblies of two well-defined graphene quantum dots of varying size: HBC, where the aromatic core consists of 42 C atoms and CQD with 78 C atoms. The synthesis of HBC and CQD were achieved using bottom-up chemical methods, while their assembly was studied using steady-state UV/Vis spectroscopy, X-ray scattering and electron microscopy. While HBC forms long ordered fibers, CQD was found not to assemble well. The exciton size and dynamics were studied using time-resolved laser spectroscopy. At early time (~100 fs), the exciton was found to delocalize over ~1-2 molecular units in both assemblies, which reflects the confined nature of excitons in carbon-based materials and is consistent with the calculated value of ~2 molecular units. Exciton-exciton annihilation measurements provided the exciton diffusion lengths of 16 nm and 3 nm for HBC and CQD, respectively.
... Fluorescence is excited at 405 nm. Figure 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 Non-radiative relaxation of the excited state is typically much faster than radiative relaxation, particularly so when energy or electron transfer occurs. On the single-molecule level, energy transfer rates have been measured at cryogenic temperatures by quantifying the spectral broadening arising from excited-state dephasing, [29][30] an approach which is not easily transferable to ambient conditions; and at room temperature using phase-coherent excitation, which is limited to a few hundred femtoseconds. 16 Figure 3 illustrates a donor-acceptor complex, in which a conjugated phenylene-butadiynylene oligomer absorbs light and then passes the excitation energy to one of two boron-dipyrromethene (bodipy) dye endcaps. ...
Article
Most measurements of fluorescence lifetimes on the single-molecule level are carried out using avalanche photon diodes (APDs). These single-photon counters are inherently slow and their response shows a strong dependence on photon energy, which can make deconvolution of the instrument response function (IRF) challenging. An ultrafast time resolution in single-molecule fluorescence is crucial, e.g., in determining donor lifetimes in donor-acceptor couples which undergo energy transfer, or in plasmonic antenna structures, where the radiative rate is enhanced. We introduce a femtosecond double-excitation (FeDEx) photon correlation technique, which measures the degree of photon antibunching as a function of time delay between two excitation pulses. In this boxcar integration, the time resolution of fluorescence transients is limited solely by the laser pulse length and is independent of the detector IRF. The versatility of the technique is demonstrated with a custom-made donor acceptor complex with one donor and two acceptors; and with single dye molecules positioned accurately between two gold nanoparticles using DNA origami. The latter structures show ~75-fold radiative-rate enhancements and fluorescence lifetimes of down to 17 ps, which is measured without the need of any deconvolution. With the potential of measuring sub-picosecond fluorescence lifetimes plasmonic antenna structures can now be optimized further.
... In natural photosynthetic systems, the bacteriochlorophylla (Bchl-a) molecules are anchored in peptide matrices [6,7]. The interaction of porphyrins with polymer systems [8][9][10][11], as well as photophysical processes in polymer systems [12][13][14][15], has received increasing attention. ...
Article
Full-text available
An amphiphilic cationic tripyridiniumylporphyrin monomer, i.e., ZnTrMPyP, was synthesized and copolymerized with acrylamide in water and dimethyl sulfoxide, respectively, to prepare the water-soluble random copolymer P-D and microblock copolymer P-W. The association behavior and fluorescence quenching between the copolymers and tetra(p-sulfonato phenyl)porphyrin (Fe(III)TSPP(Cl)) were studied via absorption and emission spectra. The results showed that relatively discrete pendant groups of ZnTrMPyP within P-D could form a ground state complex with FeTSPP by electrostatic interactions, and both static and dynamic mechanisms were active in this quenching process. In contrast, the microblock porphyrin pendant groups within P-W interacted with FeTSPP as an entity, and static quenching was dominant in this process. Salt effects on the formation of the copolymers and FeTSPP complex were also investigated through the addition of KNO3, and the results showed that this association can be weakened by the electrostatic shielding effect, and the fluorescence quenching constant could be reduced.
... Photosynthesis is the basis for all life activities, in which the sunlight is absorbed as light energy and then transformed into chemical energy in numerous chloroplast pigments through light-harvesting systems [1]. Subsequently, the excitation energy constantly transfers from chlorophyll to chlorophyll, and finally converts into chemical energy at the reaction center [2][3][4][5][6][7]. A large amount of antenna around the reaction center is a prominent feature of the light-harvesting system. ...
Article
In the present work, we have designed and synthesized an amphiphilic pyridinium-functionalized anthracene (AP) compound, which can self-assemble into nanoparticles in aqueous solution. Moreover, AP can form AP-CB7 and AP-CB8 complexes with cucurbit[7]uril (CB7) and cucurbit[8]uril (CB8) through host-guest interactions in aqueous solution, in which the pyridinium moiety or the pyridinium and anthracene moieties were encapsulated into the cavity of CB7 or CB8. AP-CB7 and AP-CB8 can further self-assemble into nanospheres and nanorods, respectively. More interestingly, three light-harvesting systems were constructed in aqueous solution by using the positively charged supramolecular assemblies of AP, AP-CB7, and AP-CB8 as an energy donor and a negatively charged fluorescent dye Eosin Y disodium salt (EY) as an energy acceptor, in which three different energy donors based on different supramolecular assemblies with various structures exhibited entirely distinguishing energy transfer performance. The energy transfer efficiency and antenna effect are 17.8% and 10.3 for AP, 19.2% and 17.6 for AP-CB7, 22.3% and 15.1 for AP-CB8 after addition of the energy acceptor EY at a donor/acceptor ratio of 100:1, respectively.
Article
Main observation and conclusion In the present work, an artificial light-harvesting system with fluorescence resonance energy transfer (FRET) is successfully fabricated in aqueous sodium dodecyl sulfonate (SDS) micellar systems. Since the tight and orderly arrangement of dodecyl in the SDS micelles is hydrophobic, tetra-(4-pyridylphenyl)ethylene (4PyTPE) can be easily encapsulated into the hydrophobic layer of SDS micelles through noncovalent interaction, which exhibits aggregation-induced emission (AIE) phenomenon and can be used as energy donor. By using amphoteric sulforhodamine 101 (SR101) fluorescent dye attached to the negatively charged surface of SDS micelles through electrostatic interaction as energy acceptor, the light-harvesting FRET process can be efficiently simulated. Through the steady-state emission spectra analysis in the micelle-mediated energy transfer from 4PyTPE to SR101, the fluorescence emission can be tuned and white light emission with CIE coordinates of (0.31, 0.29) can be successfully achieved by tuning the donor/acceptor ratio. More importantly, to better mimic natural photosynthesis, the SDS micelles with 4PyTPE and SR101 FRET system showed enhanced catalytic activity in photochemical catalysis for dehalogenation of α-bromoacetophenone in aqueous solution and the photocatalytic reaction could be extended to gram levels. This article is protected by copyright. All rights reserved.
Article
The highly efficient excitation energy transfer (EET) processes in photosynthetic light‐harvesting complexes have attracted much recent research interests. Experimentally, spectroscopic studies have provided important information on the energetics and EET dynamics. Theoretically, due to the large number of degrees of freedom and the complex interaction between the pigments and the protein environment, it is impossible to simulate the whole system quantum mechanically. Effective Hamiltonian models are often used, in which the most important degrees of freedom are treated explicitly and all the other degrees of freedom are treated as a thermal bath. However, even with such simplifications, solving the real‐time quantum dynamics could still be a difficult task. A particular challenging case in simulating the EET dynamics and related spectroscopic phenomena lies in the so‐called intermediate coupling regime, where the intermolecular electronic couplings and the electronic–vibrational couplings are of similar strength. In this article, we review theoretical studies of linear and nonlinear spectroscopic signals of photosynthetic light‐harvesting complexes, using the nonperturbative hierarchical equations of motion (HEOM) approach. Simulations were performed for the EET dynamics, various types of linear spectra, two‐dimensional electronic spectra, and pump–probe spectra. Benchmark tests of several approximate methods related to the HEOM approach were also discussed. The results show that the nonperturbative HEOM approach is an effective method in simulating the EET dynamics and spectroscopic signals of photosynthetic light‐harvesting complexes. Important insights into EET pathways, quantum effects including quantum delocalization, and quantum coherence in photosynthetic light‐harvesting complexes were also obtained through such simulations. This article is categorized under: • Theoretical and Physical Chemistry > Reaction Dynamics and Kinetics • Theoretical and Physical Chemistry > Spectroscopy • Software > Simulation Methods
Article
To conduct rapid microscope observations with the excitation spectral measurement for photosynthetic organisms, a wavelength-dispersive line-focus microscope was developed. In the developed system, fluorescence signals at multiple positions on a sample excited with different wavelengths can be detected as a two-dimensional image on the EMCCD camera at the same time. Using the developed system, one can obtain excitation spectra at every pixel over the excitation wavelength range from 635 to 695 nm, which covers the full range of the Qy bands of both chlorophyll-a and chlorophyll-b. Recording the reference laser spectra at the same time ensures robust measurement against the moderate spectral fluctuation in the excitation laser. Using an objective lens with a numerical aperture of 0.9, the lateral and axial resolutions of 0.56 and 1.08 μm, respectively, were achieved. The theoretically limited and experimentally estimated spectral resolutions of the excitation spectral measurement were 0.86 and 1.3 nm, respectively. The validity of the system was demonstrated by measuring fluorescent beads and single cells of a model alga, Chlamydomonas reinhardtii. Intrachloroplast inhomogeneity in the relative intensity of the chlorophyll-b band could be visualized in Chlamydomonas cells. The inhomogeneity reflects the intrachloroplast variation in the local peripheral antenna size.
  • M Z Papiz
M. Z. Papiz et al., Trends Plant Sci. 1, 198 (1996).
  • G Mcdermott
G. McDermott et al., Nature 374, 517 (1995).
  • J T M Kennis
  • A M Streltsov
  • H Permentier
  • T J Aartsma
  • J Amesz
J. T. M. Kennis, A. M. Streltsov, H. Permentier, T. J. Aartsma, J. Amesz, J. Phys. Chem. B 101, 8369 (1997).
  • R Monshouwer
  • I Ortiz De Zarate
  • F Van Mourik
  • R Van Grondelle
R. Monshouwer, I. Ortiz de Zarate, F. van Mourik, R. van Grondelle, Chem. Phys. Lett. 246, 341 (1995).
  • H.-M Wu
H.-M. Wu et al., J. Phys. Chem. 100, 12022 (1996).
  • M Chachisvilis
  • O Kühn
  • T Pullerits
  • V Sundström
M. Chachisvilis, O. Kühn, T. Pullerits, V. Sundström, J. Phys. Chem. B 101, 7275 (1997).
  • R Jimenez
  • S R Dikshit
  • S E Bradforth
  • G R Fleming
R. Jimenez, S. R. Dikshit, S. E. Bradforth, G. R. Fleming, J. Phys. Chem. 100, 6825 (1996).
  • S I E Vulto
  • J T M Kennis
  • A M Streltsov
  • J Amesz
  • T J Aartsma
S. I. E. Vulto, J. T. M. Kennis, A. M. Streltsov, J. Amesz, T. J. Aartsma, J. Phys. Chem. B 103, 878 (1999).
  • V Nagarajan
  • W W Parson
V. Nagarajan and W. W. Parson, Biochemistry 36, 2300 (1997).
  • M A Bopp
  • Y Jia
  • L Li
  • R J Cogdell
  • R M Hochstrasser
M. A. Bopp, Y. Jia, L. Li, R. J. Cogdell, R. M. Hochstrasser, Proc. Natl. Acad. Sci. U.S.A. 94, 10630 (1997).