ISSN 0006-2979, Biochemistry (Moscow), 2023, Vol. 88, No. 10, pp. 1528-1543 © The Author(s) 2023. This article is an open access publication.
Russian Text © The Author(s), 2023, published in Biokhimiya, 2023, Vol. 88, No. 10, pp. 1847-1866.
1528
REVIEW
Similarities and Differences in Photochemistry
of Type I and Type II Rhodopsins
Mikhail A. Ostrovsky
1,2
, Olga A. Smitienko
2
, Anastasia V. Bochenkova
3
,
and Tatiana B. Feldman
1,2,a
*
1
Faculty of Biology, Lomonosov Moscow State University, 119991 Moscow, Russia
2
Emanuel Institute of Biochemical Physics, 119334 Moscow, Russia
3
Faculty of Chemistry, Lomonosov Moscow State University, 119991 Moscow, Russia
a
e-mail: feldmantb@mail.ru
Received July 1, 2023
Revised July 20, 2023
Accepted August 12, 2023
Abstract The diversity of the retinal-containing proteins (rhodopsins) in nature is extremely large. Fundamental similar-
ity of the structure and photochemical properties unites them into one family. However, there is still a debate about the
origin of retinal-containing proteins: divergent or convergent evolution? In this review, based on the results of our own
and literature data, a comparative analysis of the similarities and differences in the photoconversion of the rhodopsin
of typesI andII iscarried out. The results of experimental studies of the forward and reverse photoreactions of bacte-
riorhodopsin (type I) and visual rhodopsin (type II) in the femto- and picosecond time scale, photo-reversible reaction
of octopus rhodopsin (type II), photovoltaic reactions, as well as quantum chemical calculations of the forward photo-
reactions of bacteriorhodopsin and visual rhodopsin are presented. The issue of probable convergent evolution of type I
and type II rhodopsins is discussed.
DOI: 10.1134/S0006297923100097
Keywords: retinal-containing proteins, visual rhodopsin, bacteriorhodopsin, convergent evolution, photochemistry, femto-
second spectroscopy, quantum chemical calculations
Abbreviations: RCP,retinal-containing protein; RPSB,retinal protonated Schiff-base; BR,bacteriorhodopsin of Halobacterium
salinarum archaeon; Rh,bovine visual rhodopsin of Bostaurus; FC,Franck–Condon state; wave packet, a set of coherent excited
vibrational states; CI,conical intersection of potential energy surfaces; HOOP,hydrogen-out-of-plane vibrations; PES,potential
energy surface
* To whom correspondence should be addressed.
INTRODUCTION
Family of the retinal-containing proteins(RCPs) in-
clude three types of rhodopsin: microbial (typeI), ani-
mal (type II), and recently discovered heliorhodopsins
(type III) [1-7]. Despite the fact that their functions are
very diverse, fundamental similarity of the structure,
7-α-helical topography of the protein part and retinal as
a chromophore group, as well as of their photochemi-
cal and spectral properties is astounding. Obviously, this
raises the question of their evolutionary origin. Bearing
in mind conflicting opinions on their divergent [8-10] or
convergent [11-15] evolution, we consider the latter more
favorable. In other words, it has been suggested that there
is no common ancestor for all three types of rhodopsins.
And, if this is true, most likely, the pressure of external
factors (Darwin’s natural selection) and physiological
need resulted in such amazing similarity of such unrelat-
ed RCPs. In this regards comparison and understanding
of evolution of each type of rhodopsins seem to be an
independent topic that attract a lot of interest among the
researchers. The ever-increasing number of publications
on this topic testifies to this[3-5,11,16-18].
It became clear recently that the diversity of RCPs
is very large. At present they have been found in all do-
mains of life bacteria, archaea, and eukaryotes, as well
PHOTOCHEMISTRY OF TYPE I AND II RHODOPSINS 1529
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
as in giant viruses. Type I rhodopsins are characteris-
tic for bacteria, archaea, viruses, and lower eukaryotes;
they have very diverse functions with main ones includ-
ing light-driven energy (ion pumps) and photo-infor-
mation (sensor rhodopsins, cation, and anion channels)
functions [6]. Type II rhodopsins are characteristic for
higher animals; in the majority of cases, they are repre-
sented by the specialized G-protein-coupled receptors,
which ensure mainly photo-information functions, pre-
dominantly vision function [1,2]. Type III rhodopsins
are widely available in live organisms from the same
domains of life as typeI rhodopsins, and they probably
perform photo-information functions [7].
Initially the term rhodopsin was referred to only pro-
teins mediating vision. The protein discovered in 1876
by Ferenz Boll was first named a vision compound
(Sehestoff), next, due to its color, “visual purple” (Seh-
purpur), and later “rhodopsin” from the Greek words
rhodo”– pink and “opsis”– sight. The retinal-contain-
ing proton pump of halophilic archaeon (Halobacterium
salinarum) discovered by Walther Stoeckenius and Dieter
Oesterhelt almost 100 years later was named bacteri-
orhodopsin by analogy with the visual rhodopsin [19].
Atpresent the name rhodopsin is applied to all mem-
bers of typeI, II, and III RCPs. Their common features
are, first, the structure of apo-protein, opsin, with seven
transmembrane α-helices, and, second, cofactor (chro-
mophore)– retinal that absorbs light quanta. It becomes
more obvious with time that the 7-α-helical protein do-
main of RCPs is both very conserved and, at the same
time, it is characterized with very high plasticity. As to
retinal, it is, as a rule, covalently bound via the pro-
tonated Schiff-base (RPSB) with the lysine amino acid
residue of opsin in the seventh transmembrane α-he-
lix (TM7), and exist as all-trans (types I and III rho-
dopsins) or 11-cis (type II rhodopsins) isomeric forms.
It should be noted that the chromophore center is the
most conserved domain of opsin. Protein environment
of RPSB is crucial for both spectral tuning of the rho-
dopsin molecule, and for facilitating ultrafast and effi-
cient photochemical reaction of the chromophore isom-
erization, which is the basis for functioning of all RCPs.
Considering the role of protein environment of RPSB
in the chromophore center of opsin, it is acceptable
to consider the process of photoisomerization as cata-
lyzed by the protein. Furthermore, the issue of the nature
of protein–chromophore interactions (steric, electro-
static, hydrogen-bonding, and hydrophobic) is actively
investigated [20].
With regards to membrane topology of the protein
part of the molecule, it was established that in types I
and II rhodopsins N-end is facing outside the cell
and C-end– inside, while the opposite is observed in
typeIII rhodopsins with N-end facing inside the cell and
C-end – outside [4, 7, 21]. The reason and biological
significance of such arrangement of typeIII rhodopsin
is not yet clear. It is also surprizing that the “inverted”
topology of the protein part of the molecule has been
also observed in the olfactory G-protein-binding recep-
tors of insects[22].
Type I rhodopsins of bacteria and archaea, includ-
ing bacteriorhodopsin that performs the simplest pho-
tosynthesis, are among the most ancient proteins in the
biosphere, they emerged in prokaryotic cells around
3.8 billion years ago. Type I rhodopsins of single-cell
eukaryotes emerged 3.2billion years ago, while type II
rhodopsins of higher animals, including visual rhodop-
sin, appeared in multi-cellular eukaryotes less than 1bil-
lion years ago [5,11,18,23-25].
It has been suggested in the framework of the the-
ory of convergent evolution of rhodopsins that the
lysosomal cysteine transporters with 7-α-helical struc-
ture were predecessors of the microbial rhodopsins[15].
The cAMP-dependent G-protein-coupled receptors
with classic 7-α-helical structure, i.e., not containing
retinal receptors, are considered as predecessors of an-
imal rhodopsins [12-14], and retinal chromophore was
inserted in the chromophore centers of opsins later.
It has been suggested in one of the studies in the
framework of the theory of divergent evolution that, first,
type II rhodopsins originated from the cAMP receptors
(same as other G-protein-coupled receptors of classA),
from which later type I rhodopsins were generated via
horizontal gene transfer from eukaryotes to prokaryotes
[9]. In another work, based on optimization of electron
properties of the chromophore without considering clear
differences in amino acid sequences, the authors suggest
existence of the divergent pathway of evolution, but in
this case from type I to type II rhodopsins [8]. In one
other study the authors suggested their common origin
based on comparison of the structures of Na-binding
centers in the conserved amino acid region of the TM6
and on the functionally significant tilt of this helix in
the microbial rhodopsins and in the G-protein-coupled
receptors [10].
With regards to the recently discovered type III
rhodopsins, their origin remains obscure. According to
the suggestion of the authors of one study, heliorho-
dopsins originated from typeII eukaryotic rhodopsins,
which were next captured by the giant viruses and were
transferred to prokaryotic cells [16]. This implies, similar
to one of the hypotheses of the origin of typeI rhodo-
psins [9], unusual direction of evolution from eukaryotes
toprokaryotes.
In any case, the whole array of the data accumu-
lated so far allows to favor the notion on convergent in-
dependent evolution of all types of RCPs. In this review
similarities and differences in molecular mechanisms of
photochemical reaction in type I and type II rhodop-
sins are considered. Based on our own experimental
data and quantum-chemical calculations wepresent our
arguments in favor of the convergent evolution of RCPs.
OSTROVSKY et al.1530
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
COMPARISON OF THE PRIMARY
PHOTOREACTIONS INTYPEI
AND TYPEII RHODOPSINS
Rhodopsin functioning is based on the photochem-
ical reaction of isomerization of the RPSB chromo-
phore group (all-trans → 13-cis in typeI rhodopsins and
11-cis → all-trans in typeII rhodopsins) (Fig. 1a) [1, 2,
26-30]. Photoreaction occurs in the excited state and is
characterized with unique parameters, which are deter-
mined by both chemical properties of the chromophore
itself and effect of protein environment on the chromo-
phore [26, 31-33]. In the process, the energy of light
quantum is stored as the energy of the highly twisted
configuration of the isomerized RPSB in the chromo-
phore center[34]. Moreover, in the case of type I rho-
dopsins additional contribution to the energy storage is
provided by the change of the hydrogen bond network in
the RPSB region[27]. These processes occur on a fem-
to- and early-picosecond time scale [27]. Next, trans-
formation of the strained chromophore conformation
into a relaxed state occurs accompanied by the release
of stored energy, which, in turn, leads to re-arrangement
of the protein environment in the chromophore center.
Thisprocess eventually initiates global structural chang-
es of the entire protein part of the molecule required for
its functioning.
Mechanisms of photoinduced reactions of type I
and type II rhodopsins are presented in this section
of the review using the most studied representatives of
these two classes of RCPs, proton pump– bacteriorho-
dopsin of the H. salinarum archaeon (BR) and G-pro-
tein coupled receptor– bovine visual rhodopsin of Bos
taurus(Rh), as examples.
Stationary absorption spectra of BR and Rh are
shown in Fig.1b, which consist of α-, β-, and γ-bands
with the latter defined by the opsin absorption at 280 nm.
α- and β-bands are associated with the absorption of
RPSB as a part of the chromophore center. Depending
on the isomeric form, protonation state of the Schiff
base, and on peculiarities of the chromophore center
structure the spectral, photochemical, and a number
of other functionally important properties of the mol-
ecule change significantly[35]. Position of the α-band
absorption determines spectral range of rhodopsin func-
tioning (300-700nm)[2]. In the case of BR, maximum
of the α-band absorption is at 568 nm [Fig. 1b (1)], and
in the case of Rh – at 498 nm [Fig. 1b (2)]. Position
of the β-ionone ring significantly affects position of
the RPSB absorption maximum [36]. In BR the chro-
mophore group has planar trans-configuration of the
C
5
= C
6
and C
7
= C
8
bonds relative to the C
6
–C
7
single
bond [6-s-trans; Fig. 1a (1)], while in Rh it has cis-con-
figuration [6-s-cis; Fig. 1a (2)], which is not planar due
to steric hindrance causing decrease of the conjugation
length in the chromophore and shift of the Rh absorp-
tion maximum towards shorter wavelengths.
Forward photoreaction of type I and type II rhodop-
sins. Absorption of a quantum of light results in photo-
chemical reaction of RPSB isomerization and formation
of the primary photoproducts with spectral properties
different from the initial state. In this review our own
results on the dynamics of the primary processes of pho-
totransformation of BR and Rh obtained using method
of femtosecond absorption laser spectroscopy [37-46]
and supplemented with the data of quantum-chemical
calculations are presented [32,47-50].
Difference spectra and kinetic curves of the pho-
toinduced absorption of BR and Rh presented in Fig. 2
are produced with probing pulse delay time of up to 10ps.
The first photoinduced signals observed in the early fem-
tosecond time scale include absorption and stimulated
emission from the S
1
excited state [Fig. 2a (2and 3);
Fig. 2b (2)]. These signals are followed by absorption
of the first photoproduct in the ground electronic state
S
0
, and bleaching appears in the region of absorption
Fig. 1. Physicochemical characteristics of the chromophore group in type I and type II rhodopsins. a) Chemical structure of the rhodopsin
chromophore group– all-trans RPSB in typeI rhodopsins(1) and 6-s-cis–11-cis RPSB in typeII rhodopsins(2). Reactive bond is shown in bold.
b)Stationary absorption spectra of the suspension of purple membranes containing bacteriorhodopsin of the Halobacterium salinarum archaeon
(BR)(1) and of the detergent extract of the visual rhodopsin of Bos taurus(Rh)(2) normalized to α-band absorption.
PHOTOCHEMISTRY OF TYPE I AND II RHODOPSINS 1531
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
Fig. 2. Forward photoreactions of BR and Rh. a)Spectra of photoinduced absorption of BR recorded with delay times: –0.2(1), 0.05(2), 0.12(3),
0.5(4), 1(5), and 10(6) ps. b) Spectra of photoinduced absorption of Rh recorded with delay times: –0.2(1), 0.03(2), 0.1(3), 0.2(4), 0.8(5),
and 10(6)ps. Stationary absorption spectra of BR(a) and Rh(b) are presented with negative sign(7). c,d)Normalized kinetic curves of pho-
toinduced absorption of BR(1) and Rh(2) recorded at the absorption band of the excited state(c) with probing wavelengths of 470(BR) and
410(Rh)nm and in the absorption band of the photoreaction products(d) with probing wavelengths of 640(BR) and 580(Rh)nm. Kinetic curves
are shown in the linear scale of delay time up to 2ps and further in the logarithmic scale. The figure is adapted from[44] with permission.
of rhodopsin in the ground state [Fig. 2a (5), 600-
700 nm; Fig. 2b (4), 540-700 nm]. Within several pico-
seconds the absorption band of the first photoproduct
shifts slightly towards shorter wavelengths, which re-
flects formation of the next product as a result of the
processes of vibrational relaxation of the chromophore
group and its amino acid environment [Fig. 2a (6);
Fig. 2b (6)]. A fraction of the excited BR and Rh mol-
ecules returns back into the initial state with non-isom-
erized RPSB, which determined quantum yield of the
reaction.
Comparative analysis of the photoinduced absorp-
tion spectra of BR and Rh demonstrates differences in the
position of absorption bands and stimulated emission, as
well as in the time of formation of the primary products.
In the case of BR, the signals of the excited state(I
460
)
appear after delay time of 100 fs (Fig. 2a(2and3); 400-
540 nm and 700-880 nm) [42-44]. Within 1ps these sig-
nals practically completely disappear and are replaced
with the positive absorption signal of the first product
(J
625
) [Fig. 2a (5)], which contains RPSB in 13-cis-con-
figuration. The next product (K
590
) is formed within a
picosecond time interval [Fig.2a(6)].
In the case of Rh, the signals from the excited
state(Rh
*
510
) appear within <30 fs (Fig. 2b (2); 410-480 nm
and 620-720 nm) [41, 42], which is significantly faster
than in the case of BR. At the delay time of 100 fs these
signals already disappear and are replaced with the pos-
itive signal of the first product (Photo
570
), which is com-
pletely formed by the time of 200 fs after absorption of
the light quantum [Fig. 2b (3and 4)] [37, 38, 41, 42].
The next product (Batho
535
) is formed within several
picoseconds [Fig.2b(5and6)].
Kinetic curves of the photoinduced absorption of
BR and Rh are presented in Fig. 2, which reflect for-
mation and decay of the excited state (I
460
and Rh
*
510
, re-
spectively; Fig. 2c) and formation of the first product of
photoreaction (J
625
and Photo
570
, respectively; Fig. 2d).
In the case of Rh, photoreaction proceeds much fast-
er (60 fs) in comparison with BR (480 fs) [40, 42-44],
which is in good agreement with the data reported by
Kochendoerfer and Mathies [51], Polli et al. [52], and
Johnson etal. [53, 54]. Time of formation of the sec-
ond product of the BR reaction (K
590
) is estimated as
1.8 ps, and of the second product of the Rh reaction
(Batho
535
)– as 2.2 ps [44]. Analysis of dynamics of the
OSTROVSKY et al.1532
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
Fig. 3. Kinetic schemes of elementary reaction of photoisomerization in type I rhodopsin– BR(a) and in typeII rhodopsin– Rh(b). Lifetimes
of the intermediate states are presented in the schemes as well as reactive(r) and non-reactive(nr) decay pathways of the excited state are marked.
Potential energy surfaces (S
0
andS
1
) of typeI rhodopsins exemplified by BR(c) and typeII rhodopsins exemplified by Rh(d) show the reaction
pathways of their forward (dark gray color) and reverse (gray color) photoreactions in femto- and pico-second time scale. FC,Franck–Condon
state; CI,conical intersection. Figure is adapted from[42] with permission.
excited state decay (both of Rh and BR) reveals that in
asmall fraction of the molecules (~4%) lifetime of the
excited state is much longer (2.4 ps), and its decay re-
sults in formation of only initial state of the RCP, hence,
this pathway has been termed non-reactive [44]. Exis-
tence of several pathways of the excited state decay, some
of which could be non-reactive, is, in general, typical
for type I rhodopsins [55-58], and in a lesser degree
for type II rhodopsins [59]. This has been associated
with the split of the reaction pathway in the Frank–
Condon state (FC) or with the initial heterogeneity of
the protein part of the molecule [28, 60]. The latter sug-
gestion has been confirmed for the sodium pump KR2
of the Krokinobacter eikastus bacterium [49, 57] and
for the proton pumps proteorhodopsin and BR [58].
The ratio of the fraction of the excited molecules un-
dergoing isomerization and the fraction of the excited
molecules returning to the initial state via the reactive
and non-reactive pathways determines quantum yield
of isomerization, which is 0.64 for BR [61] and 0.65
for Rh [62]. Kinetic scheme and lifetimes of the ob-
served processes are presented in Fig. 3,a andb.
One of distinctive features of the photoreaction of
RPSB in rhodopsins and in solution is existence of the
vibrational component in its dynamics in the early stag-
es [29, 42, 53, 63, 64]. Phases and amplitudes of these
vibrations allowed to interpret their appearance as a re-
sult of non-stationary oscillatory motion in the excited
and ground states of the reagent and reaction products.
Contrary to the classic notions, formation of the prod-
uct in this type of photochemical reactions occurs via
coordinated motion of the molecule nuclei and ends
faster than the processes of vibrational relaxation and
vibrational dephasing in the excited state [65,66]. Such
reactions are termed coherent and could be described
as a motion of the wave packet (set of coherent excited
vibrational states) first down the S
1
potential energy sur-
face(PES) and next down the S
0
PES along the reaction
coordinate. Such fast S
1
→ S
0
transition is possible due
to existence of the multidimensional conical intersection
region(CI) of S
1
/S
0
PESs, owing to which effective and
ultrafast transformation of light energy into chemical en-
ergy occurs [31,67].
The main differences in the dynamics of forward
photoreactions of BR and Rh are associated with the
structures of PESs of the S
0
, S
1
, and S
2
states of these
rhodopsins (Fig. 3, c and d), which is determined by
both isomeric form of RPSB, and by the effect of specif-
ic protein environment of the chromophore.
In the case of BR, similar to the other typeI rho-
dopsin, the forward photoreaction is described by the
3-state model including S
0
, S
1
, and S
2
states [26, 28,
55, 68], which postulates existence of a small barri-
er on the decay pathway of the S
1
state due to interac-
tion with the S
2
PES, which significantly affects dy-
namics of the reaction (Fig. 3c). Excitation of the BR
results in formation of the FC state on the S
1
PES,
which, as a result of the wave packet motion along
PHOTOCHEMISTRY OF TYPE I AND II RHODOPSINS 1533
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
Fig. 4. Optimized structures of reagents and primary products of BR and Rh– all-trans RPSB in BR
568
(a), 13-cis RPSB in K
590
(b), 11-cis RPSB in
Rh
498
(c) and all-trans RPSB in Batho
535
(d). Dihedral angles are shown by black lines with corresponding values.
all-symmetrical (C=C and C–C) vibrational modes of
RPSB, is transferred into the excited state I
460
within
40 fs. Next, the wave packet moves along the reactive hy-
drogen out-of-plane (HOOP) and torsional vibrational
modes and by overcoming a small barrier on the S
1
PES
reaches the CI region within the characteristic time of
480 fs. Inthe CI region the wave packed is divided into
two subpackets, with one of them is transferred to the
S
0
PES of the J
625
product, and the second one to the
S
0
PES of the initial state BR
568
. As a result of overcoming
the barrier on the S
1
PES the wave packet loses its coher-
ent properties, which is obvious from the absence of clear-
ly pronounced oscillations in the time-resolved signals
of the product J
625
absorption [Fig. 2d(1)] [29, 42-44,
55,56]. This type of dynamics is termed diffusive[26].
In the case of Rh, same as for other typeII rhodop-
sins, forward photoreaction is described by the 2-state
model (S
0
andS
1
) (Fig. 3d) [2, 26, 28, 69]. As a result
of excitation of the Rh molecule a FC state is formed
on the S
1
PES, which is transformed within less than
30 fs into the excited state Rh
*
510
as a result of motion
of the wave packet along the all-symmetric (C=C and
C–C) vibrational modes of RPSB, similarly to the case
of BR. Next, the wave packet moves along the out-of-
plane (HOOP and torsional) vibrational modes, reaches
the CI region within ≈60 fs, where it is divided into two
subpackets. The first subpacket moves along the S
0
PES
of Photo
570
, which is accompanied by the vibrational
relaxation of this product visualized by the clearly pro-
nounced oscillations of these kinetic curves (Fig. 2d),
while the second subpacket is transferred to the S
0
PES
of the initial state Rh
498
continuing its motion. Hence,
the S
1
→ S
0
transition of the wave packet in the process
of photoreaction of Rh occurs very fast, without devia-
tions and barriers thus preserving the significant portion
of coherence [37, 38, 42, 63]. Such type of dynamics is
termed ballistic or impulsive [26,70].
In order to explain differences in the dynamics of
photoinduced processes and specificity of the reaction
of photoisomerization in BR and Rh, quantum-chem-
ical calculations of the structures of these RCPs were
performed, and activity of the vibrational modes
during S
0
→ S
1
excitation and their link with the reac-
tive modes during non-radiative transition S
1
→ S
0
were
analyzed.
Optimized structures of the all-trans and 11-cis
RPSB in the ground electronic state in the protein envi-
ronment of BR and Rh, respectively, obtained with the
help of combined method of quantum and molecular
mechanics (QM/MM) are presented in Fig. 4 [48].
In the protein environment the chromophores
are found to have reactive bonds significantly twisted
(Fig. 4, a and c). On the other hand, these twists are
absent outside of the protein and the π-conjugated sys-
tem is planar with exception of the double bond in the
β-ionone ring. This bond, depending on the RPSB con-
figuration, is either in the conjugation plane of all other
five double bonds [6-s-trans; Fig. 1a (1)], or is out of
the conjugation plane [6-s-cis; Fig. 1a (2)]. It should be
noted that in the protein environment the out-of-conju-
gation-plane angle of the double bond in the β-ionone
ring in BR practically does not change, while in Rh it
increases significantly [48].
Hence, the protein environment indeed promotes
change of conformation and ‘twisting’ of the chromo-
phore group, which is due to its interaction with the
closest amino acid residues of the chromophore cen-
ter. This is mainly mediated by formation of hydrogen
bond between the protonated Schiff base and primary
counterion, which affects significantly torsion angle of
the reactive bond. Furthermore, attention should be also
paid to steric interaction in the area of β-ionone ring
of retinal. As expected, geometry of the primary pho-
toproducts was found to be even more twisted around
the reactive bonds due to their inability to relax in the
protein environment, which, on picosecond time scale,
remains more optimal for the initial isomers of RPSB
(Fig.4,bandd).
OSTROVSKY et al.1534
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
Fig. 5. Activity of the vibrational modes of RPSB during S
0
→S
1
transition in gas phase and in protein environment: all-trans RPSB in gas phase(a)
and inBR(b); 11-cis RPSB in gas phase(c) and inRh(d).
Interestingly enough, rotation barrier around the
single bond C
6
–C
7
does not exceed 3 kcal/mol in the gas
phase [50], which facilitates practically free transition
between the 6-s-trans and 6-s-cis conformations in the
RPSB. As was mentioned above, length of the π-con-
jugation chain is an important factor affecting position
of the absorption maximum wavelength. Intramolecular
rotation around the single C
6
–C
7
bond results in the ab-
normally wide absorption spectrum of the RPSB isomers
in the gas phase[71], while protein environment could
easily affect RPSB conformation in the region of β-ion-
one ring, and, depending on the value of dihedral angle
of the C
6
–C
7
bond, could control the length of π-con-
jugation in the chromophore thus adjusting its absorp-
tion to the certain range. Such mechanism of regulation
of photophysical properties of RPSB could be a basis for
spectral tuning in typeII rhodopsins. Noteworthy, posi-
tion of the β-ionone ring in BR does not change neither
in comparison with the gas phase, nor during transition
into the primary photoproducts, while its conformation
changes significantly in Rh. Hence, conformation of the
β-ionone ring is less significant in typeI rhodopsins in
comparison with type II rhodopsins.
Effects of twisting of the reactive bonds in different
RPSB isomers on the dynamics of isomerization in the
protein environment could be also illustrated by com-
paring it with the dynamics of photoresponse in the gas
phase. As was mentioned above, the RPSB isomers in
the isolated state have planar structure of the main part
of the π-conjugated system. It was shown with the help
of femtosecond spectroscopy and quantum-chemical
calculations that the relaxation dynamics of the electron-
ic excited state S
1
of the 11-cis RPSB occurs within the
sub-picosecond timeframe (400 fs) [32], which is com-
parable with the ultrafast times of photoisomerization
of the chromophore in Rh (50-100 fs) [40, 42, 51-54].
At the same time, specificity of the reaction and average
lifetimes of the excited state of the all-trans RPSB differ
significantly in the isolated state (3 ps)[32] and in the
protein environment in BR (≈500 fs) [29, 55, 56, 58].
It is important to mention that photoisomerization of
the planar all-trans-isomer of RPSB occurs slowly in
thegas phase, but selectively. The lowest barrier of isom-
erization in the S
1
state has been observed in the case
of the C
11
=C
12
bond, but not in the case of the C
13
=C
14
bond as in BR.
Hence, the effect of protein environment is a key
factor in the case of type I rhodopsins. While in the case
of type II rhodopsins, more advanced (from the point of
view of photochemical properties) 11-cis-isomer of RPSB
is used. Moreover, the reaction in Rh proceeds ultrafast,
which also implies significance of the effect of protein
environment of the chromophore center. Thisdifference
explains the clearly visible coherent mode of the primary
photochemical reaction in typeII rhodopsins.
The effect of conformation of the reactive bonds
in RPSB on the dynamics of photoisomerization has
been independently confirmed during investigation of
the chemically modified all-trans-chromophore twisted
around one of the double bonds. Change of the RPSB
conformation at the reactive bond results in the signif-
icant reduction of the photoisomerization barrier and
sub-picosecond times, which is typical for the reaction in
the protein environment[47].
Twisting of the reactive bond in the electronic
state S
0
not only changes the PES topography in the
state S
1
by decreasing the barrier of photoisomerization,
but also facilitates simultaneous excitation of certain
vibrational modes on transition S
0
→ S
1
[48, 49]. Inthe
process, excitation of both valent vibrations, localized
predominantly on the reactive double bond, and of
HOOP-vibrations at this bond occur. BR protein envi-
ronment in this case facilitates excitation of vibrations
of the C
13
=C
14
bond, while in the case of Rh, protein
environment facilitates excitation of the C
11
=C
12
bond.
Activity of the vibrational modes during photoexcitation
PHOTOCHEMISTRY OF TYPE I AND II RHODOPSINS 1535
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
could be estimated from the shift of minima of the PES
of the S
0
and S
1
states along these modes. The shifts
characterize changes of geometry of the chromophore in
the S
1
-state in comparison with the S
0
state. The larger is
the shift along certain vibration mode the more active is
vibration upon photoexcitation. The so-called Huang–
Rhys factor associated with the square of the shift along
the normal mode in a harmonic approximation is a use-
ful dimensionless parameter allowing evaluation of the
activity of the vibrational mode in electronic transition.
Values of the Huang–Rhys factor for the all-trans and
11-cis RPSB in the protein environment and in the gas
phase, calculated with the help of QM/MM method
using multiconfigurational quasidegenerate perturba-
tion theory of the second order, are presented in Fig. 5
[48, 49].
Two high-frequency valent vibrations of the bonds
C
11
=C
12
and C
13
=C
14
are active in the planar all-trans
and 11-cis RPSB upon excitation to the S
1
state in the
gas phase. In the case of BR, activity of the vibration-
al mode localized at the C
13
=C
14
bond becomes signifi-
cantly higher in comparison with the gas phase, while
the valent vibration of the C
11
=C
12
bond disappears en-
tirely. On the contrary, in the case of Rh, intensity of the
mode associated with vibrations of the C
11
=C
12
bond in-
creases significantly, while intensity of the C
13
=C
14
mode
decreases. Moreover, HOOP at the C
11
=C
12
bond in Rh
also become active.
Hence, increase of twisting of the reactive bond
promotes increase of activity of the corresponding va-
lent vibration during photoexcitation. It should be men-
tioned that specificity of the photo-response of RPSB
and activity of the vibrations facilitating photoisomeri-
zation at the certain double bond (in this case valent and
HOOP vibrations) are more characteristic exactly for Rh.
This is in agreement with very high rate of the photo-
chemical reaction in this type II rhodopsin. The possi-
bility of excitation of various type of reactive modes in
the same phase, which promotes coherency of the reac-
tion in Rh, is also an important factor.
Hence, the results of quantum-chemical calcula-
tions provided explanation to the experimentally ob-
served difference in the dynamics of the photoinduced
processes in BR and Rh. The main factor affecting
these processes in both cases is protein environment of
the RPSB. Photoresponse of the chromophore group
in type I and type II rhodopsins becomes highly spe-
cific in comparison with the gas phase already at the
early times of the reaction resulting in excitation of cer-
tain vibrational modes, which facilitate isomerization
of the reactive bond. The 11-cis RPSB chromophore
in Rh is faster and more selective in comparison with
the all-trans RPSB in BR. Moreover, protein environ-
ment in Rh facilitates excitation of not only valent vibra-
tions of the reactive bond, but also of the out-of-plane
vibrations of hydrogen atoms at this bond, which directly
leads to the reaction of photoisomerization. It must be
also mentioned that conformation of the β-ionone ring
in the 6-s-cis-11-cis RPSB isomer, most likely, is an im-
portant factor in regulation of photophysical properties
of the chromophore group in typeII rhodopsins, which
is associated with the mechanism of spectral tuning.
It could be concluded that, most likely, the evo-
lutionary “younger” typeII rhodopsins have more ad-
vanced chromophore center in the opsin and isomeric
form of the chromophore, which allow more efficient
forward photochemical reaction.
Photochromism of type I and type II rhodopsins.
Rhodopsins exhibit photochromic properties [30, 34, 61,
72], i.e., have ability to realize phototransitions from the
intermediate products of the forward photoreaction back
to the initial state. This ability realized in nature in a
number of rhodopsins starting with the nanosecond time
scale serves for performing certain physiological func-
tions in such type I rhodopsins as sensory rhodopsin I
of H. salinarum [73], sensory rhodopsin of cyanobacte-
ria Anabaena sp. [29], and channel rhodopsin of the sin-
gle-cell alga Chlamydomonas reinhardtii[74], and in such
typeII rhodopsins as visual rhodopsins of invertebrates
[75-77] and closely related to them melanopsins of ver-
tebrates [78].
Despite the fact that at first stages of rhodopsin
phototransformation reverse phototransitions are practi-
cally not existing in nature, investigation of such ultrafast
photochromism under invitro conditions could provide
additional information on molecular mechanisms of the
photochemical reaction of these RCPs. In this review
the results of our studies investigating reverse photore-
actions of BR and Rh initiated in femto- and early pico-
second timeframes are presented [39,41,42].
Spectrum of photoinduced absorption of BR re-
corded with the delay time of 100 ps after the first pump
pulse with wavelength of 560 nm (pulse 1) that consists
of an absorption band of the product K
590
and bleach-
ing band of the initial state BR
568
is presented in Fig. 6
[Fig. 6a (1)]. The second pump pulse with wavelength
of 680 nm (pulse 2) and delay time of 5 ps excites K
590
molecules, part of which returns to BR
568
as a result of
reverse photoreaction (13-cis → all-trans RPSB). In the
process, dark formation of K
590
decreases and absorp-
tion of BR
568
increases [Fig. 6a (2)]. Quantum yield of
the reverse reaction of BR
568
formation from the prod-
uct K
590
is 0.81 [16], which is in agreement with the re-
sults reported by Kimetal.[61] and Balashovetal. [79].
Reverse phototransition could occur not only from the
product K
590
[Fig. 6a(inset:3and5 ps)], but also from
the product J
625
[Fig. 6a (inset:1 ps)] with approximately
equal efficiency [42].
In the case of Rh action of the pulse 2 (620 nm) that
follows with delay time after pulse 1 (500 nm) of 0.2-
3.8 ps also induces reverse photoreaction [Fig.6b][39,
41, 42, 46]. Depending on the pulse 2 delay the reverse
OSTROVSKY et al.1536
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
Fig. 6. Reverse photoreactions of BR and Rh. a,b)Spectra of photoinduced absorption of BR(a) and Rh(b) recorded with delay time of 100ps
after the action of one(1) and two(2) pump pulses with delay of the second pump pulse of 5 ps(a) and200 fs(b). In the spectral regions of
pump pulses experimental curves were completed using modeling (dashed lines). Inset: values of photoinduced absorption of BR(a) andRh(b)
recorded at probing wavelengths of 635nm(a) and560nm(b) before and after the second pump pulse depending on its delay time. c,d)Kinetic
curves of photoinduced absorption of BR(c) and Rh(d) recorded after the first pump pulse at probing wavelengths of 680 nm(c) and 620 nm(d).
Figure is adapted from[41,42] with permission.
phototransition occurs (all-trans → 11-cis RPSB) from
the Photo
570
product [Fig. 6b (inset:0.2-2 ps)] or from
the Batho
535
product [Fig. 6b (inset: 3-3.8 ps)]. Ef-
ficiency of the reverse photoreaction of Rh from the
Photo
570
product depends of the oscillation phase of
the time-resolved absorption signals of this product
[Fig. 6b(inset); Fig. 6d], and time of this photoreaction
is comparable or less than the time of forward photo-
reaction [41]. Quantum yield of the reverse photore-
action of Rh from the Batho
535
product is 0.15 [41, 42,
46], which is significantly less than the quantum yield
value (0.5) at 77 K reported previously by Suzuki and
Callender[80].
The reverse photochemical process in BR and Rh
initiated at early stages could be described as a motion of
the wave packet along the right branch of the S
1
PES of
the product of initial photochemical reaction towards the
CI region of the S
1
/S
0
PES (Fig. 3,c andd). Theoretical
calculations for Rh demonstrated that this is the same
CI region, which participates in the forward photoreac-
tion [81]. Transition of the molecules from the excited
state S
1
to the S
0
PES via the CI region results in forma-
tion of the same states as in the forward reaction: BR
568
and J
625
– in the case of BR; Rh
498
and Photo
570
– in the
case of Rh. While efficiency of the reverse photoreaction
of BR practically does not depend on the delay of the
pulse 2 [Fig. 6a(inset)], for Rh it correlates directly with
the dynamics of the wave packet in thePhoto
570
product
[Fig. 6b (inset); Fig. 6d], which is associated with clearly
pronounced coherence of the forward photoreaction of
Rh. In this case, efficiency of the reverse photoreaction
correlates with the number of molecules of the Photo
570
product excited by the pulse2.
The rate of reverse photochemical reaction in rho-
dopsins is strongly affected by the particular isomeric
form of RPSB. Phototransition from trans- to cis-form
usually is slower than phototransition cistrans, and
this is true for protein environment, gas phase, as well
as solution [26, 29, 32, 52, 57, 58, 82]. This difference is
due to the structures of S
0
, S
1
, and S
2
PES, as described
above using BR and Rh as examples [Fig. 3,c andd].
It was shown by Gaietal. [56] that, in the case of BR,
during excitation of the K
590
product the reverse photo-
chemical reaction from 13-cis- to all-trans-form of RPSB
occurs within the time period of ~100 fs, which is sig-
nificantly faster that the time of forward photoreaction
(480fs). It can be suggested that the reverse dependence
would be observed for Rh. However, as has been shown
experimentally for the Photo
570
product [41] and theoret-
ically for the Batho
535
product [81], time of the reverse
phototransition from all-trans- to 11-cis-form RPSB is
comparable with the time of forward phototransition
(60 fs) or less. It could be suggested that the rate (and
selectivity) of the reverse photochemical reaction of BR
and Rh is affected by the significant twist of the reactive
bond in the chromophore in their primary products by
the angle from –39° to –20° and from –36° to –34.8°,
respectively [31, 83]. This allows fast and barrier-less
PHOTOCHEMISTRY OF TYPE I AND II RHODOPSINS 1537
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
Fig. 7. Photochromic reaction of octopus rhodopsin. a)Difference spectra of two consecutive cycles of octopus rhodopsin transformations demon-
strating reversible conversion of rhodopsin into acidic metarhodopsin: metarhodopsin minus rhodopsin (1and3, after illumination with blue
light) and photoregeneration of rhodopsin: rhodopsin minus metarhodopsin (2and4, after illumination with red light). b)Photoreversibility
of octopus rhodopsin recorded at wavelength of maximum absorption of acidic metarhodopsin (528nm) under illumination with blue light(1),
red light(2), and in the dark(3). Figure is adapted from[76] with permission.
transition to the CI region even in the case of transcis
isomerization as in the reverse photoreaction of Rh.
It must be emphasized that while quantum yields of
the forward photoreaction in BR and Rh are practically
the same (≈0.65), quantum yields of their reverse photo-
reactions differ significantly (0.81 and 0.15, respectively).
According to the theoretical calculations [31, 81], quan-
tum yield of the product formation both in the forward
and reverse photoreactions of rhodopsins is determined
by the ratio of phases of HOOP and torsional vibration
modes of RPSB at the moment of the S
1
→ S
0
transition.
Initial population of the C=C and HOOP vibrational
modes at the reactive bond on excitation also plays an
important role in efficiency of the photoreaction, as was
mentioned above. In methanol solution the 13-cis → all-
trans and all-trans → 11-cis RPSB photoreactions proceed
with very close quantum yields (0.11 and 0.14, respec-
tively)[84]. Based on this, it could be suggested that the
significant difference in the quantum yields of the analo-
gous reverse reactions in BR and Rh is due to the effects
of protein environment at the chromophore center.
It could be mentioned that, from the function-
al point of view, the low quantum yield of the reverse
reaction in Rh in comparison with BR is an advantage
facilitating more reliable realization of the forward re-
action initiating the process of phototransduction. Light
absorption by type I rhodopsins serves primarily the en-
ergy conversion function, as in the case of BR, and re-
verse photoreactions likely cannot significantly affect
their efficiency. However, in the case of type II rho-
dopsins, which are G-protein-coupled receptors, 11-cis
RPSB acts as an effective ligand antagonist maintain-
ing low thermal “dark” noise of the photoreceptor cell.
On absorption of light, the isomerized all-trans RPSB
becomes a powerful ligand-agonist, and the activated Rh
initiates the process of phototransduction. Inthis case
the possibility of emergence of the reverse photoreaction
could significantly reduce efficiency of initiation of the
phototransduction process.
Hence, lower efficiency of the reverse photoreac-
tion of Rh, which increases reliability of the forward
photoreaction, could be considered as one of the ar-
guments in favor of selection in the course of evolution
of the 11-cis-isomer as a chromophore group in type II
rhodopsins. This indicates more advanced structure of
the chromophore center in Rh and evolutionary selec-
tion of such chromophore (11-cis RPSB), which allows
efficient phototransduction process. Obviously, under
real physiological conditions, the probability of the sec-
ond light quantum hitting the so short-lived intermedi-
ate of Rh photolysis as Batho
535
and more so Photo
570
is extremely low, however, potential possibility of reduc-
ing the probability of reverse photoreaction in the struc-
ture of Rh chromophore center exists.
At the same time, physiological significance of the
reverse photochemical reactions of type II rhodopsins
ininvertebrate animals from the late intermediate stag-
es of phototransformation is obvious. In the majority
of invertebrate animals this is one of the main ways for
regeneration (photoregeneration) of the visual pigment
[75, 77]. In particular, final product of the forward pho-
toreaction of the octopus rhodopsin is the so-called
acidic metarhodopsin, and of the reverse reaction
again rhodopsin (Fig.7a) [76].
In the course of forward photoreaction, several in-
termediates with different absorption spectra are formed,
and in the course of reverse photoreaction– there are
only two states corresponding to conformational chang-
es first in the vicinity of chromophore, and, next, in
the entire protein, which return rhodopsin into initial
state. As we have demonstrated with the detergent ex-
tracts of the octopus rhodopsin, the photoreversible re-
action of typeII rhodopsin in invertebrates is extremely
stable, and under invitro conditions it can be repeated
multiple times for a significantly long period of time
(Fig.7b)[76].
Comparison of mechanisms of phototransformation
of type I and type II rhodopsins. Comparison of the
OSTROVSKY et al.1538
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
primary photoreactions of BR and Rh reveals both com-
mon and distinguishing features. Model kinetic schemes
(Fig.3,a andb) have much in common, and reflect the
suggestion that heterogeneity of the excited states of the
investigated rhodopsins is associated with heterogeneity
of their initial states. In the course of BR and Rh reac-
tions intermediates are formed that have similar proper-
ties: I
460
, J
625
, K
590
, and Rh
*
510
, Photo
570
, Batho
535
, respec-
tively. Elementary act of isomerization of RPSB proceeds
via transition through the CI of the S
1
/S
0
PES, and time
of this transition is in femtosecond time scale. The first
products of photoinduced transformations of BR and Rh
with isomerized RPSB (J
625
and Photo
570
, respectively)
are formed in the ground(S
0
) state, and are vibrational-
ly excited precursors of the following products (K
590
and
Batho
535
, respectively) formed in the timeframe of sev-
eral picoseconds. At this stage isomerization of RPSB is
completed, and part of the light quantum energy is pre-
served in the strained structures of the products K
590
and
Batho
535
. Quantum yields of the forward photoreaction
of BR and Rh are practically the same (~0.65). Reaction
of photoisomerization of RPSB in BR and Rh, same as
in other rhodopsins, as well as in solution, is coherent
due to the inherent properties of the chromophore.
Dynamics and efficiency of the RPSB photoreac-
tion in rhodopsins depends on its isomeric form (through
the length of conjugated π-bonds) and on interactions
with the protein environment, which results not only in
the additional twisting of the chromophore, but also in
the strong electrostatic interactions with the counteri-
on. These factors affect the structure of the S
0
, S
1
, and
S
2
PES of the chromophore, by, for example, decreas-
ing barrier in the S
1
PES of type I rhodopsins, which
significantly affects rate and coherence of the reaction.
Activity of certain vibrational modes also changes during
electronic transition S
0
→ S
1
. Selective and simultaneous
excitation of the vibrational modes associated with the
reactive modes of the non-radiative transition S
1
→ S
0
into the reaction products leads at the early stages of
photoinduced dynamics to high speed and high quantum
yield of the reaction of photoisomerization.
It can be concluded that interaction of RPSB with
opsin significantly increases the rate and quantum yield
of the photoreaction in comparison with the parameters
observed in gas phase and in solution, and selectivity of
the reaction reaches 100%. The finest regulation of these
parameter is achieved in type II rhodopsins, in which
the reaction has more pronounced coherence. The chro-
mophore–protein interactions also could be the cause
of heterogeneity of the initial state of RCPs, because
it is not observed in the gas phase [32].
Decrease of the length of π-conjugation in the
6-s-cis–11-cis RPSB in comparison with the all-trans-
isomer also leads to the shift of absorption maximum
of rhodopsin towards blue region of the spectrum and,
which is very important, to increase of the barrier of
thermal 11-cistrans isomerization [26]. The first ef-
fect is actively used by typeII rhodopsins as one of the
factors of spectral tuning. The second factor is crucial
for reduction of the ‘dark’ noise of the photoreceptor
cell, which is also much facilitated by the limited volume
of the chromophore center [85]. At the same time, in
the case of BR and some other typeI rhodopsins, the
chromophore center is rather spacious and does not pre-
vent thermal trans → 13-cis isomerization occurring in
the dark.
All abovementioned features of the 11-cis RPSB are
extremely important for functioning of type II rhodop-
sins that perform photo-information function. This is,
likely, the reason why the 11-cis RPSB was select-
ed in the course of evolution as a chromophore group
of visual rhodopsins in both invertebrate and verte -
brate animals.
It should be emphasized that although the reaction
of photoisomerization is common for both type I and
type II rhodopsins, the final stages of the photoactivat-
ed processes differ significantly. In type I rhodopsins,
the 13-cis RPSB is subjected at one of the final stages
to thermal isomerization into the all-trans-form thus
closing the photoactivated processes into a photocy-
cle (Fig. 8a). In type II rhodopsins of the majority of
invertebrates, absorption of the second light quantum
Fig. 8. Schemes of type I rhodopsins photocycle exemplified by BR (a) and of photolysis of type II rhodopsins of vertebrates exemplified
by Rh(b).
PHOTOCHEMISTRY OF TYPE I AND II RHODOPSINS 1539
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
by one of the long-lived intermediate products of pho-
totransformation is required to close the cycle. The cy-
cle is not closed in type II rhodopsins of the vertebrates,
and the isomerized all-trans-retinal is released from the
chromophore center of the opsin, which is called pho-
tobleaching or rhodopsin photolysis (Fig.8b). In other
words, type II rhodopsin molecule of vertebrate ani-
mals– is a single-use molecule. For further function-
ing– rhodopsin regeneration– a new retinal molecule
in 11-cis conformation should be delivered to the chro-
mophore center (see details in the review by Ostrovsky
and Feldman[20]).
General similarities of the primary photochem-
ical reactions of type I and type II rhodopsins and
their significant differences in the following stages of
phototransformation have been manifested clearly in
their photoelectric reactions. In the end of 1970s and
beginning of 1980s using the method developed by
L. A. Drachev [86], we in collaborations with Drachev
and the Skulachev group compared in detail genera-
tion of photopotential for Rh on a photoreceptor disk
membrane from rods and for BR on a purple membrane
of haloarchaea [87-89]. As expected, both transition
all-trans → 13-cis in BR, and transition 11-cis → all-trans
in Rh were accompanied by the emergence of fast pho-
toelectric response. However, there was a significant
difference: while potential generation on the purple
membrane was accompanied also by the proton trans-
fer, which is essential for the photoenergetic function
of BR, in the case of Rh such transfer was absent. Itis
exactly the moment when physiological pathways of
type I and type II rhodopsins diverge. Absence of the
proton transfer across the photoreceptor disk mem-
brane was demonstrated experimentally in our later
study[90].
In the case of Rh, conformational changes in the
protein part of the molecule occur simultaneously with
the charge separation, which first results, at the me-
tarhodopsin II stage, in rapid deprotonation of RPSB
followed by slow release of the proton into external me-
dium[91]. Asa result of these conformational rearrange-
ments, Rh at the stage of metarhodopsinII acquires abil-
ity to interact with G-protein (transducin) and to initiate
the process of phototransduction [24, 92]. Based on its
characteristic times and polarity, the rapid component of
the Rh photoresponse is, in essence, the so-called early
receptor potential, which is observed during regular elec-
trophysiological separation from the retina [93].
While in the case of BR, with its photoactivated
processes connected into a cycle, stable electric po-
tential is generated on the purple membrane during
all period of illumination, in the case of Rh, which is
undergoing photolysis, rapid photoelectric response
irreversibly drops long before the light is switched
off, and there is no response to the repeated illu-
mination [88,89].
CONCLUSIONS
Hence, it can be concluded that the similarities of
type I and type II rhodopsins are associated with the
nature of their common chromophore group, retinal,
which, to a large extent, determines the mechanism
of photoreaction as a key stage of RCP functioning.
And the observed differences in the nature of photore-
actions of typeI and typeII rhodopsins most likely are
explained by the different isomeric form of the chromo-
phore selected in the course of evolution in accordance
with their functions. It can be suggested in the frame-
work of convergent evolution of type I and type II rho-
dopsins that the similar and at the same time plastic
7-α-helical protein structures of these two types evolved
independently driven by the need to optimize electronic
properties of their chromophore groups in the all-trans
and 11-cis isomeric forms.
Similarity of the structure and mechanism of ac-
tions of such complex molecular system as retinal-con-
taining transmembrane light-sensitive proteins is due
tothe pressure of natural selection with (must be em-
phasized) certain physicochemical limitations. The se-
lection itself was pursuing the most efficient function of
each type of rhodopsins.
It is absolutely clear that the mechanisms of evo-
lution of all three currently known types of RCPs [mi-
crobial (typeI), animal (typeII), and recently discov-
ered heliorhodopsins (typeIII)] require further detailed
investigation.
Contributions. M.A.O. concept and supervision of
the study; M.A.O., O.A.S., A.V.B., and T.B.F. writing
and editing of the paper.
Funding. This work was financially supported by the
Ministry of Science and Higher Education of the Rus-
sian Federation (agreement no.075-15-2020-795, inner
no.13.1902.21.0027).
Acknowledgments. The authors are grateful to
Dr. P. A. Kusochek and V. V. Logvinov for performing
part of quantum-chemical calculations. The calculations
were carried out using the HPC resources of the Lab-
oratory of Quantum Photodynamics (RSC Tornado)
provided through the Lomonosov Moscow State Uni-
versity Program of Development.
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.
Open access. This article is licensed under a Cre-
ative Commons Attribution 4.0 International License,
which permits use, sharing, adaptation, distribution, and
reproduction in any medium or format, as long as you
give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons license,
and indicate if changes were made. The images or other
OSTROVSKY et al.1540
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
third-party material in this article are included in the ar-
ticle’s Creative Commons license, unless indicated oth-
erwise in a credit line to the material. If material is not
included in the article’s Creative Commons license and
your intended use is not permitted by statutory regula-
tion or exceeds the permitted use, you will need to obtain
permission directly from the copyright holder. To view
a copy of this license, visit http://creativecommons.org/
licenses/by/4.0/.
REFERENCES
1. Spudich, J. L., Yang, C.-S., Jung, K.-H., and Spudich,
E.N. (2000) Retinylidene proteins: structures and func-
tions from archaea to humans, Annu. Rev. Cell Dev. Biol.,
16, 365-392, doi:10.1146/annurev.cellbio.16.1.365.
2. Ernst, O. P., Lodowski, D. T., Elstner, M., Hegemann,P.,
Brown, L.S., and Kandori, H. (2014) Microbial and an-
imal rhodopsins: structures, functions, and molecular
mechanisms, Chem. Rev., 114, 126-163, doi: 10.1021/
cr4003769.
3. Nagata, T., and Inoue, K. (2021) Rhodopsins at a glance,
J.Cell Sci., 134, jcs258989, doi:10.1242/jcs.258989.
4. Rozenberg, A., Inoue, K., Kandori, H., and Béjà, O.
(2021) Microbial rhodopsins: The last two decades,
Annu. Rev. Microbiol., 75, 427-447, doi:10.1146/annurev-
micro-031721-020452.
5. Gordeliy, V., Kovalev, K., Bamberg, E., Rodriguez-
Valera, F., Zinovev, E., Zabelskii, D., Alekseev, A.,
Rosselli, R., Gushchin, I., and Okhrimenko, I. (2022)
Microbial rhodopsins, in Rhodopsin (Gordeliy, V., ed.)
Methods Mol. Biol., 2501, 1-52, Humana, New York, NY,
doi:10.1007/978-1-0716-2329-9_1.
6. Kandori, H. (2020) Biophysics of rhodopsins and opto-
genetics, Biophys. Rev., 12, 355-361, doi:10.1007/s12551-
020-00645-0.
7. Pushkarev, A., Inoue, K., Larom, S., Flores-Uribe, J.,
Singh, M., Konno, M., Tomida, S., Ito, S., Nakamu-
ra,R., Tsunoda, S.P., Philosof, A., Sharon, I., Yutin, N.,
Koonin, E.V., Kandori, H., and Béjà, O. (2018) A dis-
tinct abundant group of microbial rhodopsins discovered
using functional metagenomics, Nature, 558, 595-599,
doi:10.1038/s41586-018-0225-9.
8. Luk, H. L., Melaccio, F., Rinaldi, S., and Olivucci, M.
(2015) Molecular bases for the selection of the chromo-
phore of animal rhodopsins, Proc. Natl. Acad. Sci. USA,
112, 15297-15302, doi:10.1073/pnas.1510262112.
9. Mackin, K. A., Roy, R. A., and Theobald, D. L. (2014)
Anempirical test of convergent evolution in rhodopsins,
Mol. Biol. Evol., 31, 85-95, doi:10.1093/molbev/mst171.
10. Shalaeva, D. N., Galperin, M. Y., and Mulkidjanian,
A. Y. (2015) Eukaryotic G protein-coupled receptors as
descendants of prokaryotic sodium-translocating rho-
dopsins, Biol. Forward, 10, 63, doi: 10.1186/s13062-
015-0091-4.
11. Kojima, K., and Sudo, Y. (2023) Convergent evolution of
animal and microbial rhodopsins, RSC Adv., 13, 5367-
5381, doi:10.1039/d2ra07073a.
12. Krishnan, A., Almen, M. S., Fredriksson, R., and Schioth,
H. B. (2012) The origin of GPCRs: identification of mam-
malian like rhodopsin, adhesion, glutamate and frizzled GP-
CRs in fungi, PLoS One, 7, e29817, doi:10.1371/journal.
pone.0029817.
13. Feuda, R., Hamilton, S. C., McInerney, J. O., and
Pisani, D. (2012) Metazoan opsin evolution reveals a
simple route to animal vision, Proc. Natl. Acad. Sci. USA,
109, 18868-18872, doi:10.1073/pnas.1204609109.
14. Feuda, R., Menon, A. K., and Göpfert, M. C. (2022)
Rethinking opsins, Mol. Biol. Evol.,
39, 3, doi: 10.1093/
molbev/msac033.
15. Zhai, Y., Heijne, W. H., Smith, D. W., and Saier, M. H.,
Jr. (2001) Homologues of archaeal rhodopsins in plants,
animals and fungi: structural and functional predica-
tions for a putative fungal chaperone protein, Biochim.
Biophys. Acta, 1511, 206-223, doi: 10.1016/s0005-2736
(00)00389-8.
16. Bulzu, P.-A., Kavagutti, V. S., Andrei, A.-S., and
Ghai, R. (2022) The evolutionary kaleidoscope of rho-
dopsins, mSystems, 7, e00405-22, doi:10.1128/msystems.
00405-22.
17. Govorunova, E. G., Sineshchekov, O. A., and Spudich,
J. L. (2022) Emerging diversity of channelrhodopsins and
their structure-function relationships, Front. Cell. Neuros-
ci., 15, 800313, doi:10.3389/fncel.2021.800313.
18. Ostrovsky, M. A. (2017) Rhodopsin: evolution and
comparative physiology, Paleontol. J., 51, 562-572,
doi:10.1134/S0031030117050069.
19. Oesterhelt, D., and Stoeckenius, W. (1971) Rhodopsin-
like protein from the purple membrane of Halobacterium
halobium, Nat. New Biol., 233, 149-152, doi:10.1038/new-
bio233149a0.
20. Ostrovsky, M. A. and Feldman, T. B. (2012) Chemistry
and molecular physiology of vision: light-sensitive protein
rhodopsin, Russ. Chem. Rev., 81, 1071-1090, doi:10.1070/
RC2012v081n11ABEH004309.
21. Kovalev, K., Volkov, D., Astashkin, R., Alekseev, A.,
Gushchin, I., Haro-Moreno, J. M., Chizhov, I.,
Siletsky, S., Mamedov, M., Rogachev, A., Balandin, T.,
Borshchevskiy, V., Popov, A., Bourenkov, G., Bam-
berg,E., Rodriguez-Valera, F., Büldt, G., and Gordeliy,V.
(2020) High-resolution structural insights into the helior-
hodopsin family, Proc. Natl. Acad. Sci. USA, 117, 4131-
4141, doi:10.1073/pnas.1915888117.
22. Benton, R., Sachse, S., Michnick, S. W., and Vosshall,
L.B. (2006) Atypical membrane topology and heteromer-
ic function of Drosophila odorant receptors invivo, PLoS
Biol., 4, e20, doi:10.1371/journal.pbio.0040020.
23. Lamb, T. D., Collin, S. P., and Pugh, E. N., Jr. (2007)
Evolution of the vertebrate eye: Opsins, photorecep-
tors, retina and eye cup, Nat. Rev. Neurosci., 8, 960-976,
doi:10.1038/nrn2283.
PHOTOCHEMISTRY OF TYPE I AND II RHODOPSINS 1541
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
24. Hofmann, K. P., and Lamb, T. D. (2023) Rhodopsin,
light-sensor of vision, Prog. Retin. Eye Res., 93, 101116,
doi:10.1016/j.preteyeres.2022.101116.
25. Rozanov, A. Yu. (2009) Conditions of life on early earth
after 4.0 billion years ago, Problems of Origin of Life,
PINRAN, Moscow, pp.185-201.
26. Gozem, S., Luk, H. L., Schapiro, I., and Olivucci, M.
(2017) Theory and simulation of the ultrafast dou-
ble-bond isomerization of biological chromophores,
Chem. Rev., 117, 13502-13565, doi:10.1021/acs.chemrev.
7b00177.
27. Kandori, H. (2011) Protein-controlled ultrafast photo-
isomerization in rhodopsin and bacteriorhodopsin,
Supramolecular Photochemistry: Controlling Photo-
chemical Processes, Chapter 14, 571-595, doi: 10.1002/
9781118095300.ch14.
28. Diller, R. (2008) Primary reactions in retinal proteins,
in Ultrashort laser pulses in biology and medicine. Bio-
logical and Medical Physics, Biomedical Engineering,
(Braun M., Gilch P., Zinth W., eds), Springer, Berlin.
Heidelberg, Germany, Chapter 10, 243-277, doi:10.1007/
978-3-540-73566-3_10.
29. Wand, A., Gdor, I., Zhu, J., Sheves, M., and Ruhman, S.
(2013) Shedding new light on retinal protein photochem-
istry, Annu. Rev. Phys. Chem., 64, 437-458, doi:10.1146/
annurev-physchem-040412-110148.
30. Hampp, N. (2000) Bacteriorhodopsin as a photochromic
retinal protein for optical memories, Chem. Rev., 100,
1755-1776, doi:10.1021/cr980072x.
31. Schapiro, I., Melaccio, F., Laricheva, E. N., and Olivuc-
ci,M. (2011) Using the computer to understand the chem-
istry of conical intersections, Photochem. Photobiol. Sci.,
10,867-886, doi:10.1039/c0pp00290a.
32. Kiefer, H.V., Gruber, E., Langeland, J., Kusochek, P.A.,
Bochenkova, A.V., and Andersen, L. H. (2019) Intrinsic
photoisomerization dynamics of protonated Schiff-base
retinal, Nat. Commun., 10, 1210, doi:10.1038/s41467-019-
09225-7.
33. Agathangelou, D., Roy, P. P., del Carmen Marin, M.,
Ferre, N., Olivucci, M., Buckup, T., Léonard, J., and
Haacke, S. (2021) Sub-picosecond C=C bond photo-
isomerization: evidence for the role of excited state mix-
ing, Comptes Rendus Physique, 22, 111-138, doi: 10.5802/
crphys.41.
34. Birge, R. R., Cooper, T. M., Lawrence, A. F., Masthay,
M.B., Vasilakis, C., Fan Zhang, C., and Zidovetzki, R.
(1989) A spectroscopic, photocalorimetric, and theoreti-
cal investigation of the quantum efficiency of the primary
event in bacteriorhodopsin, J.Am. Chem. Soc., 111, 4063-
4074, doi:10.1021/ja00193a044.
35. Wang, W., Geiger, J. H., and Borhan, B. (2013) The pho-
tochemical determinants of color vision, BioEssays, 36,
65-74, doi:10.1002/bies.201300094.
36. Okada, T., Sugihara, M., Bondar, A.-N., Elstner, M.,
Entel, P., and Buss, V. (2004) The retinal conformation
and its environment in rhodopsin in light of a new 2.2 Å
crystal structure, J.Mol. Biol., 342, 571-583, doi:10.1016/
j.jmb.2004.07.044.
37. Smitienko, O. A., Shelaev, I. V., Gostev, F. E., Feldman,
T.B., Nadtochenko, V.A., Sarkisov, O.M., and Ostro-
vsky, M. A. (2008) Coherent processes in formation of
primary products of rhodopsin photolysis, Dokl. Biochem.
Biophys., 421, 194-198.
38. Smitienko, O. A., Mozgovaya, M. N., Shelaev, I. V.,
Gostev, F. E., Feldman, T. B., Nadtochenko, V.A., Sark-
isov, O. M., and Ostrovsky, M. A. (2010) Femtosecond
formation dynamics of primary photoproducts of visual
pigment rhodopsin, Biochemistry (Moscow), 75, 25-35,
doi:10.1134/s0006297910010049.
39. Mozgovaya, M. N., Smitienko, O. A., Shelaev, I. V.,
Gostev, F. E., Feldman, T. B., Nadtochenko, V. A.,
Sarkisov, O. M., and Ostrovsky, M.A. (2010) Photochr-
omism of visual pigment rhodopsin on the femtosecond
time scale: coherent control of retinal chromophore isom-
erization, Dokl. Biochem. Biophys., 435, 302-306.
40. Nadtochenko, V. A., Smitienko, O. A., Feldman, T. B.,
Mozgovaya, M. N., Shelaev, I. V., Gostev, F. E., Sarkisov,
O. M., and Ostrovsky, M. A. (2012) Conical intersection
participation in femtosecond dynamics of visual pigment
rhodopsin chromophore cis-trans photoisomerization,
Dokl. Biochem. Biophys., 446, 242-246.
41. Smitienko, O., Nadtochenko, V., Feldman, T., Balatska-
ya, M., Shelaev, I., Gostev, F., Sarkisov, O., and Ostro-
vsky, M. (2014) Femtosecond laser spectroscopy of the
rhodopsin photochromic reaction: A concept for ultra-
fast optical molecular switch creation (ultrafast reversible
photoreaction of rhodopsin), Molecules, 19, 18351-18366,
doi:10.3390/molecules191118351.
42. Feldman, T. B., Smitienko, O. A., Shelaev, I. V., Gostev,
F. E., Nekrasova, O. V., Dolgikh, D. A., Nadtochenko,
V. A., Kirpichnikov, M. P., and Ostrovsky, M. A. (2016)
Femtosecond spectroscopic study of photochromic re-
actions of bacteriorhodopsin and visual rhodopsin,
J. Photochem. Photobiol. B, 164, 296-305, doi: 10.1016/
j.jphotobiol.2016.09.041.
43. Smitienko, O. A., Nekrasova, O. V., Kudryavtsev, A.V.,
Yakovleva, M. A., Shelaev, I. V., Gostev, F. E., Dol-
gikh, D.A., Kolchugina, I. B., Nadtochenko, V. A., Kir-
pichnikov, M. P., Feldman, T. B., and Ostrovsky, M. A.
(2017) Femtosecond and picosecond dynamics of re-
combinant bacteriorhodopsin primary reactions com-
pared to the native protein in trimeric and monomeric
forms, Biochemistry (Moscow), 82, 490-500, doi:10.1134/
S0006297917040113.
44. Smitienko, O., Feldman, T., Petrovskaya, L., Nekraso-
va,O., Yakovleva, M., Shelaev, I., Gostev, F., Cherepan-
ov, D., Kolchugina, I., Dolgikh, D., Nadtochenko, V.,
Kirpichnikov, M., and Ostrovsky, M. (2021) Comparative
femtosecond spectroscopy of primary photoreactions of
Exiguobacterium sibiricum rhodopsin and Halobacteri-
um salinarum bacteriorhodopsin, J.Phys. Chem.B, 125,
995-1008, doi:10.1021/acs.jpcb.0c07763.
OSTROVSKY et al.1542
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
45. Medvedeva, A. S., Smitienko, O. A., Feldman, T. B.,
and Ostrovsky, M. A. (2020) Comparative study of pho-
tochemistry of microbial rhodopsins (type I) and ani-
mal’s rhodopsins (Type II), Zh. Evol. Biokhim. Fiziol., 56,
519-523, doi:10.31857/S0044452920070943.
46. Ostrovsky, M. A., and Nadtochenko, V. A. (2021) Fem-
tochemistry of rhodopsins, Russ. J. Phys. Chem. B,
15, 344-351, doi:10.1134/S1990793121020226.
47. Gruber, E., Kabylda, A. M., Nielsen, M. B., Rasmussen,
A. P., Teiwes, R., Kusochek, P. A., Bochenkova, A. V.,
and Andersen, L. H. (2022) Light driven ultrafast bioin-
spired molecular motors: Steering and accelerating pho-
toisomerization dynamics of retinal, J. Am. Chem. Soc.,
144, 69-73, doi:10.1021/jacs.1c10752.
48. Kusochek, P. A., Logvinov, V. V., and Bochenkova, A.V.
(2021) Role of the protein environment in photoisom-
erization of type I and type II rhodopsins: a theoreti-
cal perspective, Moscow Univ. Chem. Bull., 76, 407-416,
doi:10.3103/S0027131421060110.
49. Kusochek, P. A., Scherbinin, A. V., and Bochenkova,
A.V. (2021) Insights into the early-time excited-state dy-
namics of structurally inhomogeneous rhodopsin KR2,
J. Phys. Chem. Lett., 12, 8664-8671, doi: 10.1021/
acs.jpclett.1c02312.
50. Toker, Y., Svendsen, A., Bochenkova, A. V., and Ander-
sen, L. H. (2012) Probing the barrier for internal rotation
of the retinal chromophore, Angew. Chem. Int. Ed. Engl.,
51, 8757-8761, doi:10.1002/anie.201203746.
51. Kochendoerfer, G. G., and Mathies, R. A. (1996) Spon-
taneous emission study of the femtosecond isomerization
dynamics of rhodopsin, J.Phys. Chem., 100, 14526-14532,
doi:10.1021/jp960509+.
52. Polli, D., Altoè, P., Weingart, O., Spillane, K. M., Man-
zoni, C., Brida, D., Tomasello, G., Orlandi, G., Kuku-
ra, P., Mathies, R. A., Garavelli, M., and Cerullo, G.
(2010) Conical intersection dynamics of the primary pho-
toisomerization event in vision, Nature, 467, 440-443,
doi:10.1038/nature09346.
53. Johnson, P. J. M., Halpin, A., Morizumi, T., Prokhoren-
ko, V. I., Ernst, O.P., and Miller, R. J. D. (2015) Local
vibrational coherences drive the primary photochem-
istry of vision, Nat. Chem., 7, 980-986, doi: 10.1038/
nchem.2398.
54. Johnson, P. J. M., Farag, M. H., Halpin, A., Morizu-
mi,T., Prokhorenko, V. I., Knoester, J., Jansen, T. L.C.,
Ernst, O. P., and Miller, R. J. D. (2017) The primary
photochemistry of vision occurs at the molecular speed
limit, J. Phys. Chem. B, 121, 4040-4047, doi: 10.1021/
acs.jpcb.7b02329.
55. Hasson, K. C., Gai, F., and Anfinrud, P. A. (1996)
The photoisomerization of retinal in bacteriorhodop-
sin: experimental evidence for a three-state model, Proc.
Natl. Acad. Sci. USA, 93, 15124-15129, doi: 10.1073/
pnas.93.26.15124.
56. Gai, F., Hasson, K. C., McDonald, J. C., and Anfinrud,
P. A. (1998) Chemical dynamics in proteins: the pho-
toisomerization of retinal in bacteriorhodopsin, Science,
279, 1886-1891, doi:10.1126/science.279.5358.1886.
57. Tahara, S., Takeuchi, S., Abe-Yoshizumi, R., Inoue, K.,
Ohtani, H., Kandori, H., and Tahara, T. (2018) Origin of
the reactive and nonreactive excited states in the primary
reaction of rhodopsins: pH dependence of femtosecond
absorption of light-driven sodium ion pump rhodop-
sinKR2, J.Phys. Chem.B, 122, 4784-4792, doi:10.1021/
acs.jpcb.8b01934.
58. Chang, C.-F., Kuramochi, H., Singh, M., Abe-Yoshizu-
mi, R., Tsukuda, T., Kandori, H., and Tahara, T. (2022)
Aunified view on varied ultrafast dynamics of the prima-
ry process in microbial rhodopsins, Angew. Chem. Int. Ed.
Engl., 61, e202111930, doi:10.1002/anie.202111930.
59. Kandori, H., Futurani, Y., Nishimura, S., Shichida, Y.,
Chosrowjan, H., Shibata, Y., and Mataga, N. (2001) Ex-
cited-state dynamics of rhodopsin probed by femtosec-
ond fluorescence spectroscopy, Chem. Phys. Lett., 334,
271-276, doi:10.1016/S0009-2614(00)01457-3.
60. Kochendoerfer, G. G., and Mathies, R. A. (1995) Ultrafast
spectroscopy of rhodopsins– photochemistry at its best!
Isr.J. Chem., 35, 211-226, doi:10.1002/ijch.199500028.
61. Kim, J. E., Tauber, M. J., and Mathies, R. A. (2001)
Wavelength dependent cis-trans isomerization in vision,
Biochemistry, 40, 13774-13778, doi:10.1021/bi0116137.
62. Govindjee, R., Balashov, S. P., and Ebrey, T. G. (1990)
Quantum efficiency of the photochemical cycle of bac-
teriorhodopsin, Biophys. J., 58, 597-608, doi: 10.1016/
S0006-3495(90)82403-6.
63. Wang, Q., Shoenlein, R. W., Peteanu, L. A., Mathies,
R. A., and Shank, C. V. (1994) Vibrationally coherent
photochemistry in the femtosecond primary event of vi-
sion, Science, 266, 422-424, doi:10.1126/science.7939680.
64. Liebel, M., Schnedermann, C., Bassolino, G., Taylor,G.,
Watts, A., and Kukura, P. (2014) Direct observation of
the coherent nuclear response after the absorption of
a photon, Phys. Rev. Lett., 112, 238301, doi: 10.1103/
PhysRevLett.112.238301.
65. Zewail, A. H. (2000) Femtochemistry: atomic-scale dy-
namics of the chemical bond, J.Phys. Chem.A, 104, 5660-
5694, doi:10.1021/jp001460h.
66. Sarkisov, O. M., and Umanskii, S. Ya. (2001) Fem-
tochemistry, Russ. Chem. Rev., 70, 449-469, doi:10.1070/
RC2001v070n06ABEH000664.
67. Klessinger, M. (1995) Conical intersections and the mech-
anism of singlet photoreactions, Angew. Chem. Int. Ed.
Engl., 34, 549-551, doi:10.1002/anie.199505491.
68. Gozem, S., Johnson, P. J. M., Halpin, A., Luk, H. L.,
Morizumi, T., Prokhorenko, V. I., Ernst, O.P., Olivuc-
ci,M., and Miller, R. J. D. (2020) Excited-state vibron-
ic dynamics of bacteriorhodopsin from two-dimensional
electronic photon echo spectroscopy and multiconfigura-
tional quantum chemistry, J.Phys. Chem. Lett., 11, 3889-
3896, doi:10.1021/acs.jpclett.0c01063.
69. Rivalta, I., Nenov, A., Weingart, O., Cerullo, G., Garavel-
li, M., and Mukamel, S. (2014) Modelling time-resolved
PHOTOCHEMISTRY OF TYPE I AND II RHODOPSINS 1543
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
two-dimensional electronic spectroscopy of the primary
photoisomerization event in rhodopsin, J.Phys. Chem.B,
118, 8396-8405, doi:10.1021/jp502538m.
70. Hayashi, S., Tajkhorshid, E., and Schulten, K. (2009)
Photochemical reaction dynamics of the primary event of
vision studied by means of a hybrid molecular simulation,
Biophys.J., 96, 403-416, doi:10.1016/j.bpj.2008.09.049.
71. Rajput, J., Rahbek, D. B., Andersen, L. H., Hirshfeld, A.,
Sheves, M., Altoè, P., Orlandi, G., and Garavelli, M.
(2010) Probing and modeling the absorption of retinal pro-
tein chromophores invacuo, Angew. Chem. Intl. Ed. Engl.,
49, 1790-1793, doi:10.1002/anie.200905061.
72. Yoshizawa, T., and Wald, G. (1963) Pre-lumirhodop-
sin and the bleaching of visual pigments, Nature, 197,
1279-1286, doi:10.1038/1971279a0.
73. Takahashi, T., Mochizuki, Y., Kamo, N., and Kobatake,Y.
(1985) Evidence that the long-lifetime photointermedi-
ate of s-rhodopsin is a receptor for negative phototaxis in
Halobacterium halobium, Biochem. Biophys. Res. Commun.,
127, 99-105, doi:10.1016/s0006-291x(85)80131-5.
74. Bruun, S., Stoeppler, D., Keidel, A., Kuhlmann, U.,
Luck, M., Diehl, A., Geiger, M. A., Woodmansee, D.,
Trauner, D., Hegemann, P., Oschkinat, H., Hilde-
brandt,P., and Stehfest, K. (2015) Light-dark adaptation
of channelrhodopsin involves photoconversion between
the all-trans and 13-cis retinal isomers, Biochemistry,
54, 5389-5400, doi:10.1021/acs.biochem.5b00597.
75. Dixon, S. F., and Cooper, A. (1987) Quantum efficien-
cies of the reversible photoreaction of octopus rhodop-
sin, Photochem. Photobiol., 46, 115-119, doi: 10.1111 /
j.1751-1097.1987.tb04744.x.
76. Ostrovsky, M. A., and Weetall, H. H. (1998) Octopus rho-
dopsin photoreversibility of a crude extract from whole
retina over several weeks’ duration, Biosens. Bioelectron.,
13, 61-65, doi:10.1016/S0956-5663(97)00078-X.
77. Inoue, K., Tsuda, M., and Terazima, M. (2007) Photore-
verse reaction dynamics of octopus rhodopsin, Biophys.J.,
92, 3643-3651, doi:10.1529/biophysj.106.101741.
78. Koyanagi, M., and Terakita, A. (2008) Gq-coupled rho-
dopsin subfamily composed of invertebrate visual pigment
and melanopsin, Photochem. Photobiol., 84, 1024-1030,
doi:10.1111/j.1751-1097.2008.00369.x.
79. Balashov, S. P., Imasheva, E. S., Govindjee, R., and
Ebrey, T. G. (1991) Quantum yield ratio of the forward
and back light reactions of bacteriorhodopsin at low
temperature and photosteady-state concentration of the
bathoproduct K, Photochem. Photobiol., 54, 955-961,
doi:10.1111/j.1751-1097.1991.tb02116.x.
80. Suzuki, T., and Callender, R. H. (1981) Primary photo-
chemistry and photoisomerization of retinal at 77 degrees
K in cattle and squid rhodopsins, Biophys.J., 34, 261-270,
doi:10.1016/S0006-3495(81)84848-5.
81. Schapiro, I., Ryazantsev, M. N., Frutos, L. M., Ferré,N.,
Lindh, R., and Olivucci, M. (2011) The ultrafast pho-
toisomerizations of rhodopsin and bathorhodopsin are
modulated by bond length alternation and HOOP driven
electronic effects, J. Am. Chem. Soc., 133, 3354-3364,
doi:10.1021/ja1056196.
82. Schoenlein, R. W., Peteanu, L. A., Mathies, R. A., and
Shank, C. V. (1991) The first step in vision: Femtosec-
ond isomerization of rhodopsin, Science, 254, 412-415,
doi:10.1126/science.1925597.
83. Altoè, P., Cembran, A., Olivucci, M., and Garavelli, M.
(2010) Aborted double bicycle-pedal isomerization with
hydrogen bond breaking is the primary event of bacteri-
orhodopsin proton pumping, Proc. Natl. Acad. Sci. USA,
107, 20172-20177, doi:10.1073/pnas.1007000107.
84. Koyama, Y., Kubo, K., Komori, M., Yasuda, H., and
Mukai, Y. (1991) Effect of protonation on the isomeriza-
tion properties of n-butylamine Schiff base of isomeric reti-
nal as revealed by direct HPLC analyses: Selection of isom-
erization pathways by retinal proteins, Photochem. Photo-
biol., 54, 433-443, doi:10.1111/j.1751-1097.1991.tb02038.x.
85. Ikuta, T., Shihoya, W., Sugiura, M., Yoshida, K., Wa-
tari, M., Tokano, T., Yamashita, K., Katayama, K.,
Tsunoda, S. P., Uchihashi, T., Kandori, H., and Nure-
ki, O. (2020) Structural insights into the mechanism of
rhodopsin phosphodiesterase, Nat. Commun., 11, 5605,
doi:10.1038/s41467-020-19376-7.
86. Drachev, L. A., Kaulen, A. D., and Skulachev, V. P. (1977)
Temporal characteristics of bacteriorhodopsin as a molec-
ular biological generator of current, Mol. Biol. (Moscow),
11, 1377-1387.
87. Bol’shakov, V. I., Drachev, L. A., Kalamkarov, G. R.,
Kaulen, A. D., and Skulachev, V. P. (1979) Community
of properties of bacterial and visual rhodopsins: light ener-
gy conversion to electric potential difference, Dokl. Akad.
Nauk SSSR, 249, 1462-1466.
88. Drachev, L. A., Kalamkarov, G. R., Kaulen, A. D.,
Ostrovsky, M. A., and Skulachev, V.P. (1981) Fast stag-
es of photoelectric processes in biological membranes.
II. Visual rhodopsin, Eur. J. Biochem., 117, 471-481,
doi:10.1111/j.1432-1033.1981.tb06362.x.
89. Drachev, L. A., Kaulen, A. D., Khitrina, L. V., and
Skulachev, V. P. (1981) Fast stages of photoelectric pro-
cesses in biological membranes. I. Bacteriorhodopsin,
Eur.J. Biochem., 117, 461-470, doi:10.1111/j.1432-1033.
1981.tb06361.x.
90. Shevchenko, T. F., Kalamkarov, G. R., and Ostrovsky,
M.A. (1987) The lack of H
+
transfer across the photore-
ceptor membrane during rhodopsin photolysis [in Rus-
sian], Sensory Sistems, 1, 117-126.
91. Ostrovsky, M. A., Fedorovich, I. B., and Poliak, S. E.
(1968) Change in pH of the medium during illumination
of the retina and of a suspension of outer segments of pho-
toreceptors, Biofizika, 13, 338-339.
92. Kalamkarov, G. R., and Ostrovsky, M. A. (2002) Mo-
lecular mechanisms of visual reception, Moscow: Nauka,
pp.1-280.
93. Brindley, G. S., and Gardner-Medwin, A. R. (1966) The
origin of the early receptor potential of the retina, J.Physi-
ol., 182, 185-194, doi:10.1113/jphysiol.1966.sp007817.