ISSN 0006-2979, Biochemistry (Moscow), 2023, Vol. 88, No. 10, pp. 1555-1570 © The Author(s) 2023. This article is an open access publication.
Published in Russian in Biokhimiya, 2023, Vol. 88, No. 10, pp. 1880-1897.
1555
REVIEW
Channelrhodopsins: From Phototaxis to Optogenetics
Elena G. Govorunova and Oleg A. Sineshchekov
a
*
a
e-mail: oleg_sinesh@yahoo.com
Received June 24, 2023
Revised July 9, 2023
Accepted July 9, 2023
AbstractChannelrhodopsins stand out among other retinal proteins because of their capacity to generate passive ionic
currents following photoactivation. Owing to that, channelrhodopsins are widely used in neuroscience and cardiology
as instruments for optogenetic manipulation of the activity of excitable cells. Photocurrents generated by channelrhodop-
sins were first discovered in the cells of green algae in the 1970s. In this review we describe this discovery and discuss
the current state of research in the field.
DOI: 10.1134/S0006297923100115
Keywords: rhodopsins, ion channels, photoreception, membrane potential, flagellate algae, neurons
Abbreviations: ChR, channel rhodopsin; EPC,early photoreceptor current; LPC,late photoreceptor current.
* To whom correspondence should be addressed.
INTRODUCTION
Bacteriorhodopsin, the focus of L. A. Drachev’s
early research [1-3], was the first retinal protein iden-
tified in a microbe rather than in mammalian tissues.
Thediscovery of bacteriorhodopsin by D. Oesterhelt and
W. Stoeckenius in the early 1970s [4] was a sensation,
and the ease of isolation of this protein in large quantities
from the source organism, haloarchaeon Halobacterium
salinarum, and its remarkable stability have made it a
favorable object of many biophysical, biochemical, and
structural studies [5-8]. It would not be an exaggeration
to say that today we know about bacteriorhodopsin more
than not only about any other retinal protein, but also
more than about any integral membrane protein, a model
of which it represents. In contrast to visual rhodopsins
that fulfil their photosensory function by activation of
an enzymatic transduction cascade, bacteriorhodopsin
is an electrogenic proton pump that transports protons
from the cytoplasm to the external medium and thus pro-
vides basic means of solar energy utilization in haloar-
chaeal cells.
When photoelectric activity of bacteriorhodopsin
was discovered, it stimulated search for similar pro-
cesses in other microorganisms, including eukaryotes.
Most motile pro- and eukaryotes respond to light with
a change in their swimming pattern [9, 10]. The first
sensory rhodopsins, SRI [11-13] and SRII [14], have
been discovered in H. salinarum. Their photoexcitation
does not lead to generation of transmembrane electrical
currents but activates an enzymatic cascade, eventual-
ly causing motor responses. In contrast to haloarchaea
that change their reversal frequency in response to light,
unicellular green flagellate algae demonstrate genuine
phototaxis, i.e., oriented movement along the direction
of the light beam [15]. Discussion of this phenomenon
with E. N. Kondratieva, a well-known expert on pho-
totroph microorganisms [16], and successful registration
of photoinduced changes in the membrane potential of
chroloplasts isolated from a higher plant [17, 18], in-
spired research on photosensing in green algae in the
Department of Physico-Chemical Biology of the Mos-
cow State University under the guidance of F. F. Litvin.
However, attempts to detect electrical stages of photo-
sensory transduction by insertion of microelectrodes
into algal cells, as used for recording electrical processes
associated with photosynthesis, have been unsuccess-
ful[19]. This goal could only be achieved by the devel-
opment of a new method for extracellular recording us-
ing a suction pipette by the corresponding author of this
review [20, 21].
Using this method, it has been shown that light in
the spectral range of phototaxis evokes transmembrane
photoreceptor potential in the region overlaying the eye-
spot – a cellular organoid made of carotenoid globules
that acts as a modulator of photoreceptor illumination
GOVORUNOVA, SINESHCHEKOV1556
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
during helical swimming of the cell. Isotonic propagation
of the depolarizing photoreceptor potential to the flagel-
lar membrane triggers a regenerative electrical response
similar to an action potential in neurons. Theselection
of the object, the green alga Haematococcus pluvialis of
a relatively large cell size and with the elastic cell wall,
was a major factor in this work’s success. Subsequently
the corresponding author of this review has shown that
both stages of this photoelectric sensory cascade could
be recorded directly in a cell suspension[22], which en-
abled its detection in many other green and cryptophyte
algae [23-27].
In contrast to bacteriorhodopsin, the main compo-
nent of purple membranes that constitute ~50% of the
cell surface, photoreceptor proteins that guide algal pho-
totaxis are present in their cells in very low concentra-
tions [28], so that their purification by biochemical meth-
ods is not feasible. Analysis of phototaxis action spectra
by K. Foster led to a hypothesis that these photorecep-
tors are retinal proteins [29]. A large body of indirect ev-
idences obtained during the 1980s and 1990s, including
restoration of phototaxis in “blind” mutants of the green
alga Chlamydomonas reinhardii deficient in carotenoid
biosynthesis upon the addition of exogenous retinal pro-
vided the first experimental validation of this hypothesis
[30-32]. Identification of Chlamydomonas photoreceptor
proteins or, rather, genes encoding them, at the molec-
ular level has been achieved only upon the development
of high-throughput methods of polynucleotide sequenc-
ing. It turned out that the genome of C. reinhardtii en-
codes not just one, but two such proteins [23, 33-35].
Remarkably, protein sequences, isomer composition of
the chromophore, and primary photochemical reac-
tions of C. reinhardtii rhodopsins turned out to be closer
to bacteriorhodopsin than to animal visual rhodopsins.
Analysis of photoreceptor currents in C. reinhardtii ge-
netic transformants with reduced amounts of rhodop-
sins showed that these proteins serve as photoreceptors
in phototaxis and the photophobic (photoshock) motile
response of this alga, although photoreceptor currents
generated by them differ in their properties [23, 28].
But only expression of C. reinhardtii rhodopsin
genes in animal cells, such as Xenopus oocytes, carried
out by G. Nagel, P. Hegemann, and E. Bamberg, has
revealed their unique properties [33, 34] (earlier this ap-
proach was successfully used also to study bacteriorho-
dopsin [36]). It tuned out that unlike the latter, C. rein-
hardtii rhodopsins passively transport cations (not only
protons, but also sodium, potassium, and, to a lesser
extent, calcium ions) across the cell membrane, i.e., act
as light-gated cation channels, the only such channels
known so far. This property was highlighted by introduc-
ing a new name, “channelrhodopsin(ChR)”, to desig-
nate these proteins, which rapidly replaced the earlier
suggested names, “Chlamydomonas sensory rhodopsins”
[23] and “archaeal-type rhodopsins” [35].
ChRs became very popular after they had been
shown to stimulate neuronal activity upon illumination
[37-39]. This technique, knows as “optogenetics”, has
revolutionized neuroscience and related fields of re-
search [40-43]. Moreover, it is anticipated that ChRs
might be used for gene therapy to cure many neurolog-
ical, psychiatric, and cardiovascular disorders [44, 45].
The prospect of vision restoration in patients with de-
generative retinal conditions looks most promising [46].
Photoactivated proton pumps similar to bacteriorho-
dopsin are also used in optogenetics as the tools for in-
hibition of neuronal activity [47], although they trans-
port less than one charge per captured photon, generate
smaller currents than ChRs and therefore require illu-
mination of longer duration and/or higher intensity to
change the membrane potential.
High-throughput sequencing of genomes and tran-
scriptomes led to identification of ChRs in several other
major eukaryotic lineages besides green algae. More-
over, it turned out that phototrophic [48] and even het-
erotrophic [49] protists possess ChRs that selectively
conduct anions. It also looks like cation selectivity in-
dependently emerged at least twice in ChR evolution,
and in one case the proteins preserved many features
of bacteriorhodopsin [50]. Finally, last year ChRs were
found that are more permeable for K
+
than for Na
+
[51],
and ChR diversity might not yet be exhausted.
In retinal ion pumps like bacteriorhodopsin, all steps
of the ion transport are tightly linked to specific photo-
chemical conversions of the photocycle monitored by
optical methods in purified proteins or membranes[52].
In contrast to this, in ChRs of different families photo-
currents correlate with different photochemical conver-
sions, and this correlation provides information on func-
tional mechanisms of the channels [50, 53-55].
In this review we will discuss methods of record-
ing ChR activity in native cells, the photoelectric sen-
sory cascade in flagellate algae, the photoreceptor role
of ChRs in phototaxis, and, briefly, their diversity and
optogenetic applications. For more detail of ChR molec-
ular mechanisms we recommend other reviews [56-59].
Also, many reviews cover the history and principles of
optogenetics [42, 60-65].
METHODS FOR REGISTRATION
OF ChR ACTIVITY IN PROTIST CELLS
Photoinduced changes in the membrane potential
in retinal rodes and cones brought about by closing cy-
clic-nucleotide-gated channels can be recorded with in-
tracellular electrodes [66]. However, the small cellular
volume of flagellate algae makes this approach not feasi-
ble for the study of ChR activity. Some green flagellates
such as H. pluvialis possess an elastic cell wall, which
makes it possible to suck their cells into a micropipette.
CHANNELRHODOPSINS 1557
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
Fig. 1. Principles of recording and examples of photoelectric signals recorded in the cells of green flagellate algae: a-c)principles of the suction
pipette technique(a), measurements in non-oriented cell suspensions (b), and measurements in pre-oriented cell suspensions(c); d-f)examples
ofphotocurrent traces recoded from an individual cell(d), a non-oriented cell suspension(e), and a pre-oriented cell suspension(f). The arrows
inpanels(a-c) show the time of the excitation flash. Designations: PC,photoreceptor current; RR,regenerative response; rec.,recorded.
This does not lead to formation of a gigaOhm seal that
is necessary for voltage clamp, but asymmetric distri-
bution of ChR molecules in the cell membrane allows
registration of their photoinduced responses [20, 21].
Thirteen years later this approach (known as the suction
pipette technique, Fig.1a) was applied to a C.reinhardtii
mutant lacking outer layers of the cell wall [67], and in
eight more years, to a Volvox carteri mutant [68]. C.rein-
hardtii is a model object of many biological studies, in
the genome of which the first channelrhodospin genes
were subsequently identified [23, 33-35].
The development of the suction pipette technique
enabled the discovery of photoinduced depolarization
of the membrane in green flagellate algae mediated by
their phototaxis receptors [20,21], which laid the foun-
dation for the development of optogenetics twenty-five
years later. However, application of the suction pipette
technique is not only labor-consuming, but also limited
to cells with the elastic cell wall, rare among flagellates.
In addition, sucking a cell into a pipette may activate its
mechanoreceptors [69] and in any case cannot be con-
sidered as a physiological condition. Asymmetric local-
ization of molecular generators of photocurrents in dif-
ferent regions of the cell membrane, revealed by using
the suction pipette technique, led to the hypothesis that
a total photocurrent of thousands of cells in a suspen-
sion can be recorded even without artificial increase of
the electrical resistance between different regions of the
membrane by the glass of the pipette.
In the first modification of this assay, a photoex-
citation flash is directed along the line connecting the
electrodes (Fig. 1b). Photoreceptor current generated
by the cells oriented with their photoreceptors toward
the light source exceeds that by the cells in the oppo-
site orientation, and the difference signal is detected by
the electrodes immersed in the cell suspension. Another
modification of this assay uses preorientation of the cells
with the light that causes phototaxis, or with gravity, di-
rected at the 90° angle to the excitation flash (Fig. 1c).
In this case the electrodes register a projection of the
photocurrent on the direction of the preorienting factor,
which allows instantaneous measurements of the orien-
tation degree [22, 70, 71].
Both modifications of this suspension assay allow
recording under fully physiological conditions and are
not limited by the cell size and the cell wall structure.
Such versatility and technical simplicity allowed us-
ing this method to study many C. reinhardtii mutants
with defects in the eyespot size [72] and photomotil-
ity [73, 74]. Moreover, the suspension assay was used
to confirm that phototaxis restoration in “blind” carot-
enoid-deficient mutants upon the addition of exogenous
GOVORUNOVA, SINESHCHEKOV1558
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
retinal indeed reflects reconstitution of functional pho-
toreceptor proteins rather than, for instance, retinal in-
corporation in the eyespot [31]. Furthermore, the sus-
pension assay enabled the discovery and characterization
of channelrhodopsin-mediated photoinduced electrical
cascades in several other green algae, including members
of the genera Spermatozopsis, Hafniomonas, Polytomella,
Mesostigma, and Platymonas, and in the phylogenetically
distant cryptophyte alga Cryptomonas [24-27]. Finally,
the suspension method was used as an indirect approach
to study gravitaxis [22] and chemotaxis [75] in C. rein-
hardtii, and provided a foundation for the development
of an express bioassay for detection of water pollution
with heavy metals and formaldehyde [76, 77]. A more
detailed technical description of both methods, the suc-
tion pipette technique, and the suspension assay, can be
found in the earlier published reviews [25, 70, 71].
Electrophysiological investigation of photorecep-
tor currents allowed characterization of photoreceptor
proteins in flagellate algae even before their genes were
cloned. Approximation of the light dependence of pho-
toreceptor current produced the value of 0.8 Å for the
product of the quantum yield and the optical cross sec-
tion, which is close to that of other retinal proteins [78].
The recovery of photoreceptor current after a saturating
flash provided an estimate of the duration of the photo-
receptor pigment’s photocycle (100 ms). Analysis of the
dependence of photoreceptor current amplitude on the
orientation of the polarization plane of the light stimu-
lus showed that the retinal chromophore is orientated in
parallel with the membrane plane [79, 80]. The action
spectra of photoreceptor current and photomotility of
H. pluvialis reveal a complex multiband structure, with
the most red-shifted band at 550 nm [20, 21]. This result
led to the conclusion that in flagellates the photorecep-
tor system might consist of several pigments acting as
an antenna for a bacteriorhodopsin-type protein. Sub-
sequently, a contribution of two photoreceptor proteins
in phototaxis has directly been shown in the model alga
C. reinhardtii [23], and even more candidates for this role
have been found in some other algae. However, all these
proteins are light-gated ion channels rather than photo-
activated ion pumps like bacteriorhodopsin [33,34].
PHOTOELECTRIC CASCADE IN PHOTOTAXIS
OF GREEN FLAGELLATE ALGAE
The photoelectric signal recorded in H. pluvialis cells
sucked into a micropipette comprises a gradual primary
photoreceptor current and a secondary regenerative re-
sponse that is driven by the primary current and develops
in the “all or nothing” manner. Comparison of the signs
of these signal components upon suction of different
parts of the cell into the pipette revealed that the prima-
ry inward photocurrent flows only across a small patch
of the plasma membrane overlaying the eyespot [20, 21,
67]. This intracellular structure originally was thought
to contain photoreceptor pigment [81], but later it was
found to play only an accessory role as a shading/reflect-
ing device [29, 82, 83].
After cloning of ChR genes, immunofluorescent
methods could be used to determine intracellular local-
ization of the encoded proteins. It has been shown that
both C. reinhardtii ChRs are mostly confined to the eye-
spot region [35, 84, 85], which is fully consistent with
the results of electrophysiological studies. Proteomic
analysis of the isolated eyespot preparations also showed
the presence of ChRs [86]. According to the current
view, ChR molecules are embedded in the plasma mem-
brane that shows specific ultrastructure in the eyespot
region [87]. In C. reinhardtii cells grown under a light-
dark cycle, immunofluorescence microscopy also shows
ChRs in the flagella and basal bodies, and their amount
in these organelles depends of the phase of the cycle
[88,89]. However, the functional state of ChRs found in
these locations is not yet known.
Flagellates rotate around their longitudinal axis
during swimming, so that the photoreceptor patch of
the plasma membrane experiences a periodic change
in illumination [9, 29, 90-94]. These conditions can be
modeled in a cell held on a micropipette by periodic il-
lumination at the 1 Hz frequency. At the light intensities
eliciting phototaxis in H. pluvialis, the beat frequency
of the cis-flagellum (the one closest to the eyespot) in-
creases, whereas that of the trans-flagellum decreases af-
ter switching the light on, and the opposite responses are
observed after switching the light off [78, 90, 95]. Such
asymmetric changes in the flagella beat frequency in a
freely swimming cell are expected to steer the cells in the
direction of light, i.e., lead to phototaxis. Photoinduced
motor responses of C. reinhardtii flagella have been stud-
ied in more detail and involve changes of not only beat
frequency, but also amplitude and velocity [96-99].
The regenerative electrical response occurs in the
flagellar membrane and therefore is described in the
literature as “the flagellar current” [32, 67, 70]. Paral-
lel registration of flagella beating and photoelectric re-
sponses in a cell held on a micropipette has shown that
the regenerative response triggers a change in the flagella
beating mode from a ciliary stroke to undulation, ob-
served during the photophobic response in a freely swim-
ming cell [20, 21, 78, 95, 100]. The photophobic response
consists of a sudden stop and/or a random change in the
swimming direction and is observed when the stimulus
intensity and/or duration exceeds a certain threshold [9,
10, 91, 101, 102].
The amplitude of the regenerative response is prac-
tically independent on the intensity of the light stim-
ulus, but the duration of the lag period from the be-
ginning of illumination till the onset of this response
shows an inverse dependence on it. The integral under
CHANNELRHODOPSINS 1559
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
Fig. 2. Properties of the photoreceptor current components in the green flagellate alga Chlamydomonas reihardtii: a)the kinetics of the early and
late photoreceptor currents, recorded from a cell suspension; b)the light dependence of the peak amplitude of photoreceptor current (right Y axis)
and an integral under the photocurrent curve (left Y axis).
the photocurrent curve till the onset of the regenerative
response is constant at any light intensity and increases
under red background illumination that hyperpolarizes
the membrane due to photosynthesis [19, 78, 79, 103].
These results show that the regenerative response is trig-
gered by translocation of a certain number of charges
across the membrane that depolarizes it by a few mV.
The strict dependence of the regenerative response
on the Ca
2+
concentration in the medium and its sensi-
tivity to blockers of Ca
2+
channels suggest that it is medi-
ated by opening of potential-gated Ca
2+
channels in the
flagellar membrane [20, 21, 67]. Analysis of the regen-
erative response amplitude in C. reinhardtii cells during
flagella regrowth after their amputation has shown that
these channels are distributed over the entire flagella
length [104]. Selection of a mutant deficient in the regen-
erative response [74] enabled cloning of the gene cav2,
encoding these channels [105]. Immunofluorescent mi-
croscopy shows that the CAV2 protein is predominantly
located in the proximal part of the flagella.
PHOTORECEPTOR CURRENT COMPONENTS
Experiments with nanosecond laser flash excitation
and high acquisition rate have revealed that photore-
ceptor current comprises two components with different
properties [25, 32, 78, 95, 106, 107]. These components
are resolved in the current traces recorded from all tested
organisms, which indicates common principles of their
photoreceptor systems and signal transduction chains.
The rise of the early photoreceptor current (EPC) is
limited only by the response time of the setup (<30 μs
for the suction pipette technique [106] and <3 μs for the
suspension assay [31]). In contrast, the late photorecep-
tor current(LPC) occurs after a delay up to several mil-
liseconds, the duration of which depends on the stimulus
intensity. At least two kinetic components can also be
resolved in the photocurrent decay. The light saturation
level of the EPC is determined only by photochemical
processes of photon absorption, whereas that of LPC is
observed at ~1000-fold lower intensities. Only LPC but
not EPC depends on temperature and red background
illumination that hyperpolarizes the membrane by ac-
tivating photosynthesis [19, 106]. Finally, LPC shows a
much stronger dependence on the Ca
2+
concentration in
the medium, as compared to EPC [25]. Because of that,
the actual kinetics of EPC can be determined by calcu-
lating the difference between the traces recorded before
and after these treatments (Fig.2a).
The maximal amplitudes of EPC and LPC mea-
sured at saturation yield the ratio <1 : 10. However,
membrane depolarization is proportional not to the max-
imal photocurrent, but to the number of charges trans-
ported across the membrane, i.e., the integral under the
photocurrent trace. Owing to slow closing of the LPC
channels, the membrane depolarization brought about
by LPC at saturation is practically equal to that pro-
duced by EPC atsaturation, although the levels of their
saturation differ 1000 times (Fig.2b).
Ca
2+
channels that mediate LPC are located in the
plasma membrane near the eyespot region, but not nec-
essarily within it. These secondary Ca
2+
channels are the
primary reason of the very high, nearly “single-quan-
tum” photosensitivity of phototaxis [108]. In contrast to
the Ca
2+
channels in the flagellar membrane, these chan-
nels are not voltage-gated, because they generate gradu-
al current (i.e., the current amplitude of which depends
GOVORUNOVA, SINESHCHEKOV1560
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
Fig. 3. A schematic depiction of photosensory transduction in phototaxis of flagellate algae. ChR1 and2, channelrhodopsins1 and2, respectively;
EPC and LPC, early and late photoreceptor currents, respectively.
on the stimulus intensity). At least 43 genes encoding
Ca
2+
channels have been found in the C. reinhardtii ge-
nome [109], but functions of only few of them are known.
Several Ca
2+
-binding proteins but no Ca
2+
channels have
been detected in the eyespot preparations [86, 110].
One of possible reasons for this failure is the insufficient
sensitivity of proteomics for detection of integral mem-
brane proteins. It is also possible that these presumable
channels are not homologous to any other known Ca
2+
channels and therefore their sequences cannot be recog-
nized as such by bioinformatic analysis.
Mechanisms by which EPC activates the second-
ary Ca
2+
channels in the eyespot region are not yet clear.
Heterotrimeric GTPases, and Ca
2+
-dependent protein
kinases and phosphatases have been detected in prepa-
ration of isolated eyespots [111-113]. However, these
proteins can participate in other sensory cascades, such
as that activated by photoexcitation of phototropin, also
found in the eyespot preparations [86, 110]. Phototro-
pin has been shown to mediate photoregulation of the
eyespot size and the cellular content of ChRs [114].
The lack of a lag period and the high level of light
saturation led to the conclusion that EPC reflects an ion
flow through rhodopsin molecules themselves [32,103].
Attempts to determine the ionic selectivity of channel-
rhodopsins in C. reinhardtii [115, 116] and V. carteri[68]
using the suction pipette technique have not been suc-
cessful, because photocurrent recording in the algae
cannot be carried out under voltage clamp, and because
of the difficulty to separate EPC and LPC in these ex-
periments. Ionic selectivity of these and tens of other
channelrhodopsins has been determined only by patch
clamp measurements upon channelrhodopsin expression
in cultured animal cells.
Identification of ChRs genes in C. reinhardtii and
the development of methods for genetic transforma-
tion of this model alga allowed verification of the role
of ChRs as photoreceptor proteins in phototaxis and the
photophobic response. Using the method of RNA inter-
ference, K.-H. Jung in the laboratory of J. L. Spudich
created transformants with a decreased content of each
of the two ChRs, initially named Chlamydomonas sen-
sory rhodopsins A and B [23,28], better known today
as ChRs 1 and 2 (ChR1 and ChR2), respectively. Re-
markably, knocking down of ChR1 led to stimulation of
ChR2 expression, which led to a further increase in the
ChR2/ChR1 ratio in the ChR1 transformant. Analysis of
photoelectric signals recorded from transformant cells in
response to a laser flash showed that kinetics of photore-
ceptor current generated by ChR1 is limited only by the
time response of the measuring system, whereas the cur-
rent generated by ChR2 appears after a light-dependent
delay. In other words, ChR1 generated current with the
kinetics of EPC, and ChR2, current with the kinetics
of LPC. Moreover, ChR1 current exhibited high light
saturation like EPC, whereas ChR2 current, low satu-
ration like LPC. The action spectra of ChR1 and ChR2
photocurrents were also different, which reflected their
different absorption spectra: the maximum of ChR1 cur-
rent was observed at 510 nm, and that of ChR2 current,
at470 nm. These results were corroborated by absorption
spectroscopy of purified proteins obtained by hetero-
logous expression of their genes [84, 117, 118]. Measure-
ments of photoorientation by photoelectric recording in
CHANNELRHODOPSINS 1561
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
cell suspensions and of photophobic response by motion
analysis showed that both ChRs mediate both photomo-
tility responses [23, 28]. A light scattering assay for pho-
toorientation in an independently created transformant
with a reduced ChR1 content confirmed photoreceptor
role of this protein [84]. The low light saturation of LPC
mostly generated by ChR2 suggest that the latter is the
primary photoreceptor in phototaxis, whereas ChR1,
mostly responsible for EPC, primarily contributes to the
photophobic response that is observed at higher light
intensities (Fig. 3). It has to be however noted that the
ratio of ChR1 and ChR2 is different in different strains
[23, 84, 119], which may influence their relative contri-
butions in photomotility responses.
The method of RNA interference allows to reduce
the amount of the encoded protein in the cell, but not to
prevent its biosynthesis completely. The latter has been
achieved first with insertional mutagenesis [85, 120],
and later with targeted disruption of ChR genes [119,
121, 122]. In a strain with the ChR1 gene disrupted by
zinc finger nuclease, the sensitivity of photoorientation
was decreased over 3 orders of magnitude, whereas dis-
ruption of ChR2 gene produced a smaller effect, and
knocking out both ChR1 and ChR2 genes practically
abolished the response [122]. When the CRISPR/Cas9
system was used for genome editing, knockout of either
ChR gene had little effect on the sensitivity of photoori-
entation, but disruption of both genes eliminated the
response [119]. The difference between these results is
most probably explained by the difference between the
amounts of ChR1 and ChR2 in the parental strains and
the degree of compensatory upregulation of one rhodop-
sin upon suppression of the other one.
The CRISPR/Cas9 genome editing method enabled
not only knocking out ChR genes, but also creating
C. reinhardtii strains carrying point mutations of each
of them[119]. All tested mutants exhibited reduction of
photoorientation sensitivity, although the cellular levels
of mutant proteins were also reduced in most cases, as
compared to control. A ChR2 mutant with reduced rel-
ative permeability for protons compared to Na
+
, as de-
termined by patch clamp analysis in model mammalian
cells, generated photoreceptor currents only in the pres-
ence of Na
+
in the medium, which confirmed the impor-
tance of proton currents for phototaxis in the wild type.
THE ROLE OF THE CYTOPLASMIC FRAGMENT
Molecules of all known ChRs consist of the rho-
dopsin domain, formed by seven transmembrane α-heli-
ces, and a cytoplasmic fragment made of an almost equal
number of amino acid residues. Heterologous expression
studies show that the rhodopsin domain is necessary and
sufficient for channel conductance, although at least in
one ChR the cytoplasmic fragment has been found to
influence the kinetics of channel current [123]. Almost
exclusively, only rhodopsin domains are used for optoge-
netic applications and functional studies in heterologous
systems. Functional analysis of the cytoplasmic fragment
has only become possible after the development of ge-
nome editing methods in C. reinhardtii cells [121]. Trun-
cation of the last 78 C-terminal amino acid residues of
ChR1 leads to displacement of the eyespot to a more an-
terior position, and the same effect is observed after dis-
ruption of the entire ChR1 gene [122]. Moreover, the at-
tachment of a fluorescent tag to the C-terminus impedes
trafficking of ChR1 to the plasma membrane. These ob-
servations confirm the earlier proposed hypothesis that
the cytoplasmic fragments are responsible for the correct
cellular localization of ChRs.
The important role of the cytoplasmic fragment in
the function of microbial rhodopsins has been demon-
strated earlier in sensory rhodopsins from the blue-green
alga Anabaena [124]. Modification of the cytoplasmic
fragment in a channelrhodopsin has been shown to in-
fluence photocurrent kinetics [123]. Phosphorylation of
amino acid residues in the cytoplasmic fragment of ChR1
and ChR2 has been detected by mass-spectrometric
analysis C. reinhardtii eyespot preparations [110]. Rapid
changes in the degree of ChR1 phosphorylation correlate
with the sign of phototaxis, i.e., swimming towards or
away from the light source [125]. Obviously, localization
of the photoreceptors, as well as changes in photocurrents
they generate determine the phototaxis sign.
Bioinformatic analysis of the cytoplasmic fragments
of ChRs from some protists (but not C. reinhardtii) has
identified various protein domains known to participate
in enzymatic cascades in other organisms [126, 127].
However, no evidence of the functional role of these do-
mains as parts of ChR fusion proteins has been obtained
so far. In addition, the cytoplasmic fragments of some
green algae channelrhodopsins exhibit SUMO (small
ubiquitin-like modifier) binding sites, and biochemi-
cal studies confirm reversible binding of this protein to
ChR1 linked to regulation of its stability in C. reinhardtii
cells [128]. Finally, immunoprecipitation shows that the
cytoplasmic fragments of ChR1 and ChR2 bind the small
GTPase CrARL11. CrARL11 belongs to the Arf family,
other members of which are known to mediate trafficking
of proteins to the cilia and flagella in animal cells [128].
ChR DIVERSITY
AND OPTOGENETIC APPLICATIONS
Recording of photoreceptor currents in green flag-
ellate algae has revealed a wide distribution of ChRs in
members of this taxonomic group [25-27]. Green algae
(chlorophytes) are the closest relatives of land plants,
which justified genome sequencing in many of their spe-
cies [129, 130] including those earlier studied by electro-
GOVORUNOVA, SINESHCHEKOV1562
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
physiological methods. Bioinformatic analysis of the data
obtained has identified >200 individual ChR sequences
in these microorganisms (their lists can be found in [127]
and [131]), but only few of them have been studied by
patch clamp recording upon heterologous expression in
model cells. The great majority of the latter are primarily
proton channels like C. reinhardtii ChRs, and some of
them also conduct Na
+
, K
+
and, to a lesser extent, Ca
2+
[26, 27, 33, 132, 133]. Collectively, these proteins are re-
ferred to as “cation channelrhodopsins” (CCRs) [58,
131, 134]. These ChRs are widely used for photostimula-
tion of neurons and other excitable cells [40-42].
All tested cryptophytes are capable of photoorien-
tation [135-137], and photoreceptor currents similar to
those in green flagellates have been recorded from one
fresh-water species [24]. The fully sequenced genome
of the marine cryptophyte Guillardia theta encodes tens
of rhodopsins, among which there are ChRs [48, 138],
proton pumps [139] and proteins that do not transport
ions[24]. At least five G. theta ChRs function as cation
channels when their genes are expressed in mammalian
cells [138, 140, 141], and many their homologs are found
in other cryptophytes [142]. According to their sequenc-
es and details of the transport mechanism, these proteins
are closer to haloarchaeal proton pumps than chlorophyte
ChRs, and therefore have been named “bacteriorhodop-
sin-like cation channelrhodopsins” (BCCRs) [50]. Their
similarity to haloarchaeal proton-pumping rhodopsins
is further corroborated by their trimeric structure [143,
144], whereas chlorophyte ChRs form dimers [145-147].
Most BCCRs generate small currents upon expression in
mammalian cells, but at least one their representative,
known as ChRmine (also known as Rhodomonas lens
cation channelrhodopsin1, RlCCR1), is a popular op-
togenetic tool for neuronal activation using green light
[148, 149].
Another group of G. theta ChRs exhibit exclusively
anion selectivity [48]. Their homologs have been found
in many other cryptophytes and are known as “anion
channelrhodopsins” (ACRs) [150, 151]. As the Cl
con-
centration in the soma of mature neurons is low, opening
of these channels leads to Cl
influx across the mem-
brane and inhibition of spiking. ACRs have successfully
been used as optogenetic inhibitors of neuronal activity
and behavior in worms, insects, fish, and mammals [48,
152-158]. In addition, they have been applied for opto-
genetic control of guard cell movements and growth di-
rection in higher plants [159, 160].
Besides cryptophytes, haptophyte [49] and even
some chlorophyte [127] algae have been shown to pos-
sess ACRs. Moreover, homologous proteins (or rather
genes that encode them) have been found in heterotroph
organisms from the infrakingdom of stramenopiles [49,
131, 161]. Most if not all of these organisms develop fla-
gella at least at one stage of their life cycle. It is plausible
that ACRs, as well as CCRs, acts as photoreceptors in
their phototaxis, but this hypothesis requires experimen-
tal verification.
Using ACRs, ~200 sequences of which are current-
ly known [127, 131], as optogenetic inhibitors is limited
by the direction of their photocurrents in recipient cells.
The Cl
concentration in the axon terminals is typically
higher than in the soma, so ACR photoexcitation depo-
larizes the membrane and may even lead to triggering
back-propagating action potentials [162-165]. These side
effects can be reduced by the addition of soma-targeting
motifs that decrease trafficking of ACRs to the terminals
[163, 165, 166], but complete elimination of this unde-
sired trafficking has not yet been possible.
That is why the discovery of ChRs more permeable
for K
+
than for Na
+
last year attracted much attention
of the optogenetic community. The first two proteins of
this functional class, known as “kailum channelrhodop-
sins” (KCRs), have been identified in the heterotrophic
stramenopile Hyphochytrium catenoides [51], and later
their homologs were found in some other organisms
[167, 168]. The results of the first experiments on inhibi-
tion of neuronal and cardiomyocyte activity using KCRs
look very promising [51, 167, 169], so these proteins are
anticipated to become the primary molecular instrument
for his purpose.
It is important to note that only one function of
channelrhodopsins is currently used in optogenetics
their direct channel activity. Their second function that
is dominant in native algal cells, generation of the late
(delayed) photoreceptor Ca
2+
current with 1,000-fold
magnification, has not yet found its practical applica-
tion, because the channels involved have not yet been
identified at the molecular level.
CONCLUSION
In the 1970s, when the studies on photoelectric ac-
tivity of bacteriorhodopsin and yet unidentified ChRs
were launched at the Moscow State University, it was im-
possible to predict the emergence of the entire new field
of biomedicine known today as optogenetics and based
on application of these proteins to manipulate the mem-
brane potential with light. At present, only one function
of phototaxis receptors (their channel activity) is used in
optogenetics. A complex and highly sensitive photosen-
sory cascade that mediates phototaxis is still awaiting its
applications. Nevertheless, this example serves as a very
convincing demonstration of the importance of funda-
mental studies for scientific and technical progress.
Contributions. E.G.G. writing the text; O.A.S. con-
ceptualization and supervision of the work, writing and
editing the text.
Funding. The work was supported by the private
funds of the authors.
CHANNELRHODOPSINS 1563
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
Acknowledgments. We thank all colleagues who con-
tributed publications cited in this review.
Ethics declarations. The authors declare no conflicts
of interest. This article does not contain description of
studies with the involvement of humans or animal subjects.
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
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. Drachev, L. A., Kaulen, A. D., Ostroumov, S. A., and
Skulachev, V. P. (1974) Electrogenesis by bacteriorho-
dopsin incorporated in a planar phospholipid membrane,
FEBS Lett., 39, 43-45, doi:10.1016/0014-5793(74)80012-8.
2. Drachev, L. A., Kaulen, A. D., and Skulachev, V. P. (1978)
Time resolution of the intermediate steps in the bacterior-
hodopsin-linked electrogenesis, FEBS Lett., 87, 161-167,
doi:10.1016/0014-5793(78)80157-4.
3. Drachev, L. A., Kaulen, A. D., Skulachev, V. P., Khitrina,
L.V., and Chekulaeva, L.N. (1981) Phases of photoelec-
trical response of bacteriorhodopsin [in Russian], Bio-
khimiia, 46, 998-1005.
4. Oesterhelt, D., and Stoeckenius, W. (1971) Rhodopsin-like
protein from the purple membrane of Halobacterium halo-
bium, Nature, 233, 149-152, doi:10.1038/newbio233149a0.
5. Balashov, S. P., and Litvin, F. F. (1985) Photochemical
Conversions of Bacteriorhodopsin, Moscow State University
Press, Moscow.
6. Kandori, H. (2000) Role of internal water molecules in
bacteriorhodopsin, Biochim. Biophys. Acta Bioenergetics,
1460, 177-191, doi:10.1016/s0005-2728(00)00138-9.
7. Lanyi, J. K. (2004) Bacteriorhodopsin, Annu. Rev. Physiol.,
66, 665-688, doi:10.1146/annurev.physiol.66.032102.150049.
8. Wickstrand, C., Nogly, P., Nango, E., Iwata, S., Stand-
fuss, J., and Neutze, R. (2019) Bacteriorhodopsin: Struc-
tural insights revealed using X-ray lasers and synchrotron
radiation, Annu. Rev. Biochem., 88, 59-83, doi: 10.1146/
annurev-biochem-013118-111327.
9. Sineshchekov, O. A., and Litvin, F. F. (1974) Phototaxis in
microorganisms, its mechanism and relation to photosyn-
thesis, Usp. Sovr. Biol., 78, 58-75.
10. Nultsch, W., and Häder, D.-P. (1988) Photomovement in
motile microorganisms – II, Photochem. Photobiol., 47,
837-869, doi:10.1111/j.1751-1097.1988.tb01668.x.
11. Spudich, E. N., and Spudich, J. L. (1982) Control of
transmembrane ion fluxes to select halorhodopsin-defi-
cient and other energy-trunsduction mutants of Halobac-
terium halobium, Proc. Natl. Acad. Sci. USA, 79, 4308-
4312, doi:10.1073/pnas.79.14.4308.
12. Bogomolni, R., and Spudich, J. L. (1982) Identification of
a third rhodopsin-like pigment in phototactic Halobacte-
rium halobium, Proc. Natl. Acad. Sci. USA, 79, 6250-6254,
doi:10.1073/pnas.79.20.6250.
13. Spudich, J. L., and Bogomolni, R. A. (1984) Mechanism
of colour discrimination by a bacterial sensory rhodopsin,
Nature, 312, 509-513, doi:10.1038/312509a0.
14. Takahashi, T., Tomioka, H., Kamo, N., and Kobatake,Y.
(1985) Aphotosystem other than PS370 also mediates the
negative phototaxis of Halobacterium halobium, FEMS
Microbiol. Lett.,
28, 161-164, doi: 10.1111/j.1574-6968.
1985.tb00784.x.
15. Feinleib, M. E. H., and Curry, G. M. (1971) Therelation-
ship between stimulus intensity and oriented phototactic
response (topotaxis) in Chlamydomonas, Physiol. Planta-
rum, 25, 346-352, doi:10.1111/j.1399-3054.1971.tb01453.x.
16. Kondratieva, E. N., and Gogotov, I. N. (1969) Production
of hydrogen by green photosynthetic bacteria (Chloro-
pseudomonas), Nature, 221, 83-84, doi:10.1038/221083a0.
17. Bulychev, A. A., Andrianov, V. K., Kurella, G. A., and
Litvin, F.F. (1971) Transmembrane potential of the chlo-
roplast and its photoinduced changes, Dokl. Akad. Nauk
SSSR, 197, 473-476.
18. Sineshchekov, O. A., Andrianov, V. K., Kurella, G. A.,
and Litvin, F.F. (1976) Bioelectric phenomena in a uni-
cellular green alga and their relation to phototaxis and
photosynthesis, Fiziol. Rast., 23, 229-237.
19. Sineshchekov, O. A., Sineshchekov, V. A., and Litvin,
F.F. (1978) Photoinduced bioelectrical responses in pho-
totaxis of a unicellular flagellated alga, Dokl. Akad. Nauk
SSSR, 239, 471-474.
20. Litvin, F. F., Sineshchekov, O. A., and Sineshchekov,
V.A. (1978) Photoreceptor electric potential in the photo-
taxis of the alga Haematococcus pluvialis, Nature, 271, 476-
478, doi:10.1038/271476a0.
21. Sineshchekov, O. A., Govorunova, E. G., Der, A.,
Keszthelyi, L., and Nultsch, W. (1992) Photoelectric re-
sponses in phototactic flagellated algae measured in cell
suspension, J.Photochem. Photobiol. B Biol., 13, 119-134,
doi:10.1016/1011-1344(92)85051-U.
22. Kreimer, G. (1994) Cell biology of phototaxis in flag-
ellate algae, Int. Rev. Cytol., 148, 229-310, doi: 10.1016/
S0074-7696(08)62409-2.
23. Sineshchekov, O. A., Jung, K.-H., and Spudich, J. L.
(2002) Two rhodopsins mediate phototaxis to low- and
high-intensity light in Chlamydomonas reinhardtii, Proc.
Natl. Acad. Sci. USA, 99, 8689-8694, doi: 10.1073/
pnas.122243399.
GOVORUNOVA, SINESHCHEKOV1564
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
24. Sineshchekov, O. A., Govorunova, E. G., Jung, K.-H.,
Zauner, S., Maier, U.-G., and Spudich, J.L. (2005) Rho-
dopsin-mediated photoreception in cryptophyte flagel-
lates, Biophys. J., 89, 4310-4319, doi: 10.1529/biophysj.
105.070920.
25. Sineshchekov, O. A., and Spudich, J. L. (2005) Sensory
rhodopsin signaling in green flagellate algae. in Hand-
book of Photosensory Receptors, Wiley-VCH, Weinheim,
pp.25-42, doi:10.1002/352760510X.ch2.
26. Govorunova, E. G., Spudich, E. N., Lane, C. E., Sinesh-
chekov, O. A., and Spudich, J. L. (2011) New channelrho-
dopsin with a red-shifted spectrum and rapid kinetics from
Mesostigma viride, mBio, 2, e00115-00111, doi: 10.1128/
mBio.00115-11.
27. Govorunova, E. G., Sineshchekov, O. A., Li, H., Janz,R.,
and Spudich, J.L. (2013) Characterization of a highly ef-
ficient blue-shifted channelrhodopsin from the marine
alga Platymonas subcordiformis, J.Biol. Chem., 288, 29911-
29922, doi:10.1074/jbc.M113.505495.
28. Govorunova, E. G., Jung, K.-W., Sineshchekov, O. A.,
and Spudich, J. L. (2004) Chlamydomonas sensory rho-
dopsinsA andB: cellular content and role in photopho-
bic responses, Biophys. J., 86, 2342-2349, doi: 10.1016/
S0006-3495(04)74291-5.
29. Foster, K.-W., and Smyth, R. D. (1980) Light anten-
nas in phototactic algae, Microbiol. Rev., 44, 572-630,
doi:10.1128/mr.44.4.572-630.1980.
30. Foster, K.-W., Saranak, J., Patel, N., Zarrilli, G.,
Okabe,M., Kline, T., and Nakanishi, K. (1984) Arho-
dopsin is the functional photoreceptor for phototaxis in
the unicelullar eukaryote Chlamydomonas, Nature, 311,
756-759, doi:10.1038/311756a0.
31. Sineshchekov, O. A., Govorunova, E. G., Der, A.,
Keszthelyi, L., and Nultsch, W. (1994) Photoinduced
electric currents in carotenoid-deficient Chlamydomo-
nas mutants reconstituted with retinal and its analogs,
Biophys. J., 66, 2073-2084, doi: 10.1016/S0006-3495
(94)81002-1.
32. Sineshchekov, O. A., and Govorunova, E. G. (1999) Rho-
dopsin-mediated photosensing in green flagellated al-
gae, Trends Plant Sci., 4, 58-63, doi:10.1016/s1360-1385
(98)01370-3.
33. Nagel, G., Ollig, D., Fuhrmann, M., Kateriya, S., Musti,
A. M., Bamberg, E., and Hegemann, P. (2002) Channel-
rhodopsin-1: a light-gated proton channel in green algae,
Science, 296, 2395-2398, doi:10.1126/science.1072068.
34. Nagel, G., Szellas, T., Huhn, W., Kateriya, S., Adeishvi-
li,N., Berthold, P., Ollig, D., Hegemann, P., and Bam-
berg, E. (2003) Channelrhodopsin-2, a directly light-gated
cation-selective membrane channel, Proc. Natl. Acad. Sci.
USA, 100, 13940-13945, doi:10.1073/pnas.1936192100.
35. Suzuki, T., Yamasaki, K., Fujita, S., Oda, K., Iseki, M.,
Yoshida, K., Watanabe, M., Daiyasu, H., Toh, H., Asa-
mizu, E., Tabata, S., Miura, K., Fukuzawa, H., Nakamu-
ra,S., and Takahashi, T. (2003) Archaeal-type rhodopsins
in Chlamydomonas: model structure and intracellular lo-
calization, Biochem. Biophys. Res. Commun., 301, 711-717,
doi:10.1016/s0006-291x(02)03079-6.
36. Nagel, G., Mockel, B., Buldt, G., and Bamberg, E.
(1995) Functional expression of bacteriorhodopsin in oo-
cytes allows direct measurement of voltage dependence
of light induced H
+
pumping, FEBS Lett., 377, 263-266,
doi:10.1016/0014-5793(95)01356-3.
37. Boyden, E. S., Zhang, F., Bamberg, E., Nagel, G., and
Deisseroth, K. (2005) Millisecond-timescale, genetically
targeted optical control of neural activity, Nat. Neurosci.,
8, 1263-1268, doi:10.1038/nn1525.
38. Nagel, G., Brauner, M., Liewald, J. F., Adeishvili, N.,
Bamberg, E., and Gottschalk, A. (2005) Light activation
of channelrhodopsin-2 in excitable cells of Caenorhabdi-
tis elegans triggers rapid behavioral responses, Curr. Biol.,
15, 2279-2284, doi:10.1016/j.cub.2005.11.032.
39. Li, X., Gutierrez, D. V., Hanson, M. G., Han, J., Mark,
M. D., Chiel, H., Hegemann, P., Landmesser, L. T., and
Herlitze, S. (2005) Fast noninvasive activation and inhibi-
tion of neural and network activity by vertebrate rhodopsin
and green algae channelrhodopsin, Proc. Natl. Acad. Sci.
USA, 102, 17816-17821, doi:10.1073/pnas.0509030102.
40. Deisseroth, K. (2011) Optogenetics, Nat. Methods, 8,
26-29, doi:10.1038/nmeth.f.324.
41. Yawo, H., Asano, T., Sakai, S., and Ishizuka, T. (2013)
Optogenetic manipulation of neural and non-neural
functions, Dev. Growth Differ., 55, 474-490, doi:10.1111/
dgd.12053.
42. Deisseroth, K. (2015) Optogenetics: 10years of microbi-
al opsins in neuroscience, Nat. Neurosci., 18, 1213-1225,
doi:10.1038/nn.4091.
43. Emiliani, V., Entcheva, E., Hedrich, R., Hegemann, P.,
Konrad, K. R., Lüscher, C., Mahn, M., Pan, Z.-H., Sims,
R. R., Vierock, J., and Yizhar, O. (2022) Optogenetics
for light control of biological systems, Nat. Rev. Methods
Primers, 2, 55, doi:10.1038/s43586-022-00136-4.
44. Chow, B. Y., and Boyden, E. S. (2013) Optogenetics and
translational medicine, Sci. Transl. Med., 5, 177ps175,
doi:10.1126/scitranslmed.3003101.
45. Sasse, P., Funken, M., Beiert, T., and Bruegmann, T.
(2019) Optogenetic termination of cardiac arrhythmia:
Mechanistic enlightenment and therapeutic application?
Front. Physiol., 10, 675, doi:10.3389/fphys.2019.00675.
46. Sahel, J. A., Boulanger-Scemama, E., Pagot, C., Arleo,A.,
Galluppi, F., Martel, J. N., Esposti, S. D., Delaux,A., de
Saint Aubert, J. B., de Montleau, C., Gutman, E., Audo, I.,
Duebel, J., Picaud, S., Dalkara, D., Blouin, L., Taiel, M.,
and Roska, B. (2021) Partial recovery of visual function in a
blind patient after optogenetic therapy, Nat. Med., 27, 1223-
1229, doi:10.1038/s41591-021-01351-4.
47. Chow, B. Y., Han, X., Dobry, A. S., Qian, X., Chuong,
A.S., Li, M., Henninger, M.A., Belfort, G.M., Lin,Y.,
Monahan, P. E., and Boyden, E. S. (2010) High-per-
formance genetically targetable optical neural silenc-
ing by light-driven proton pumps, Nature, 463, 98-102,
doi:10.1038/nature08652.
CHANNELRHODOPSINS 1565
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
48. Govorunova, E. G., Sineshchekov, O. A., Liu, X.,
Janz, R., and Spudich, J. L. (2015) Natural light-gated
anion channels: Afamily of microbial rhodopsins for ad-
vanced optogenetics, Science, 349, 647-650, doi:10.1126/
science.aaa7484.
49. Govorunova, E. G., Sineshchekov, O. A., Li, H.,
Wang,Y., Brown, L. S., and Spudich, J.L. (2020) Rub-
yACRs, non-algal anion channelrhodopsins with highly
red-shifted absorption, Proc. Natl. Acad. Sci. USA, 117,
22833-22840, doi:10.1073/pnas.2005981117.
50. Sineshchekov, O. A., Govorunova, E. G., Li, H., and
Spudich, J. L. (2017) Bacteriorhodopsin-like channelrho-
dopsins: Alternative mechanism for control of cation con-
ductance, Proc. Natl. Acad. Sci. USA, 114, E9512-E9519,
doi:10.1073/pnas.1710702114.
51. Govorunova, E. G., Gou, Y., Sineshchekov, O. A., Li, H.,
Lu, X., Wang, Y., Brown, L. S., St-Pierre, F., Xue, M.,
and Spudich, J. L. (2022) Kalium channelrhodopsins
are natural light-gated potassium channels that medi-
ate optogenetic inhibition, Nat. Neurosci., 25, 967-974,
doi:10.1038/s41593-022-01094-6.
52. Der, A., and Keszthelyi, L. (2001) Charge motion during
the photocycle of bacteriorhodopsin, Biochemistry (Mos-
cow), 66, 1234-1248, doi:10.1023/A:1013179101782.
53. Sineshchekov, O. A., Govorunova, E. G., Wang, J.,
Li,H., and Spudich, J. L. (2013) Intramolecular proton
transfer in channelrhodopsins, Biophys.J., 104, 807-817,
doi:10.1016/j.bpj.2013.01.002.
54. Sineshchekov, O. A., Govorunova, E. G., Li, H., and
Spudich, J.L. (2015) Gating mechanisms of a natural an-
ion channelrhodopsin, Proc. Natl. Acad. Sci. USA, 112,
14236-14241, doi:10.1073/pnas.1513602112.
55. Sineshchekov, O. A., Govorunova, E. G., Li, H.,
Wang,Y., and Spudich, J.L. (2023) Sequential absorp-
tion of two photons creates a bistable form of RubyACR
responsible for its strong desensitization, Proc. Natl.
Acad. Sci. USA, 120, e2301521120, doi: 10.1073/pnas.
2301521120.
56. Schneider, F., Grimm, C., and Hegemann, P. (2015) Bio-
physics of channelrhodopsin, Annu. Rev. Biophys., 44,
167-186, doi:10.1146/annurev-biophys-060414-034014.
57. Deisseroth, K., and Hegemann, P. (2017) Theform and
function of channelrhodopsin, Science, 357, eaan5544,
doi:10.1126/science.aan5544.
58. Govorunova, E. G., Sineshchekov, O. A., Li, H., and
Spudich, J. L. (2017) Microbial rhodopsins: Diversity,
mechanisms, and optogenetic applications, Annu. Rev.
Biochem., 86, 845-872, doi: 10.1146/annurev-biochem-
101910-144233.
59. Govorunova, E. G., Sineshchekov, O. A., and Spudich,
J. L. (2021) Emerging diversity of channelrhodopsins and
their structure-function relationships, Front. Cell Neuros-
ci., 15, 800313, doi:10.3389/fncel.2021.800313.
60. Boyden, E. S. (2011) A history of optogenetics: the devel-
opment of tools for controlling brain circuits with light,
F1000 Biol. Rep., 3, 11, doi:10.3410/B3-11.
61. Fenno, L., Yizhar, O., and Deisseroth, K. (2011) The de-
velopment and application of optogenetics, Annu. Rev.
Neurosci.,
34, 389-412, doi: 10.1146/annurev-neuro-
061010-113817.
62. Hegemann, P., and Nagel, G. (2013) From channel-
rhodopsins to optogenetics, EMBO Mol. Med., 5, 1-4,
doi:10.1002/emmm.201202387.
63. Boyden, E. S. (2015) Optogenetics and the future of neu-
roscience, Nat. Neurosci., 18, 1200-1201, doi: 10.1038/
nn.4094.
64. Richards, R., and Dempski, R. E. (2015) From photo-
taxis to biomedical applications: Investigating the molec-
ular mechanism of channelrhodopsins, in Electrophysi-
ology of Unconventional Channels and Pores (Delcour,
A.H., ed) Springer International Publishing Switzerland,
pp.361-381, doi:10.1007/978-3-319-20149-8_15.
65. Bamberg, E., Gärtner, W., and Trauner, D. (2018)
Introduction: Optogenetics and photopharmacology,
Chem. Rev., 118, 10627-10628, doi:10.1021/acs.chemrev.
8b00483.
66. Baylor, D. A., and Hodgkin, A. L. (1973) Detection
and resolution of visual stimuli by turtle photorecep-
tors, J.Physiol., 234, 163-198, doi:10.1113/jphysiol.1973.
sp010340.
67. Harz, H., and Hegemann, P. (1991) Rhodopsin-regu-
lated calcium currents in Chlamydomonas, Nature, 351,
489-491, doi:10.1038/351489a0.
68. Braun, F. J., and Hegemann, P. (1999) Two light-acti-
vated conductances in the eye of the green alga Volvox
carteri, Biophys. J., 76, 1668-1678, doi: 10.1016/S0006-
3495(99)77326-1.
69. Yoshimura, K. (1996) A novel type of mechanorecep-
tion by the flagella of Chlamydomonas, J.Exp. Biol., 199,
295-302, doi:10.1242/jeb.199.2.295.
70. Sineshchekov, O. A., and Govorunova, E. G. (2001) Elec-
trical events in photomovements of green flagellated algae,
in Comprehensive Series in Photosciences (Hader, D.-P.,
and Lebert, M., eds) Elsevier, Amsterdam, pp.245-280,
doi:10.1016/S1568-461X(01)80013-2.
71. Sineshchekov, O. A., Govorunova, E. G., and Spudich,
J. L. (2021) Probing channelrhodopsin electrical activi-
ty in algal cell populations, in Channelrhodopsin (Demp-
ski, R., ed.) vol. 2191, Humana, New York. pp. 85-96,
doi:978-1-0716-0830-2_6.
72. Kreimer, G., Overländer, C., Sineshchekov, O. A.,
Stolzis,H., Nultsch, W., and Melkonian, M. (1992) Func-
tional analysis of the eyespot in Chlamydomonas reinhard-
tii mutant ey 627, mt
, Planta, 188, 513-521, doi:10.1007/
BF00197043.
73. Pazour, G., Sineshchekov, O., and Witman, G. B. (1995)
Mutational analysis of the phototransduction pathway of
Chlamydomonas reinhardtii, J. Cell Biol., 131, 427-440,
doi:10.1083/jcb.131.2.427.
74. Matsuda, A., Yoshimura, K., Sineshchekov, O., Hiro-
no, M., and Kamiya, R. (1998) Isolation and character-
ization of novel Chlamydomonas mutants that display
GOVORUNOVA, SINESHCHEKOV1566
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
phototaxis but not photophobic response, Cell Motil. Cy-
toskeleton, 41, 353-362, doi: 10.1002/(SICI)1097-0169
(1998)41:4<353::AID-CM7>3.0.CO;2-0.
75. Govorunova, E. G., and Sineshchekov, O. A. (2003) In-
tegration of photo- and chemosensory signaling pathways
in Chlamydomonas, Planta, 216, 535-540, doi: 10.1007/
s00425-002-0901-7.
76. Govorunova, E. G., Altschuler, I. M., Häder, D. P., and
Sineshchekov, O. A. (2000) Anovel express bioassay for
detecting toxic substances in water by recording rhodop-
sin-mediated photoelectric responses in Chlamydomo-
nas cell suspensions, Photochem. Photobiol., 72, 320-326,
doi:10.1562/0031-8655(2000)072<0320:anebfd>2.0.co;2.
77. Govorunova, E. G., and Sineshchekov, O. A. (2017) Ex-
press detection of water pollutants by photoelectric re-
cording from algal cell suspensions, in Bioassays: Advanced
Methods and Applications (Häder, D.-P., and Erzinger,
G. S., eds) Elsevier, Amsterdam, pp.289-308, doi:10.1016/
B978-0-12-811861-0.00014-0.
78. Sineshchekov, O. A. (1991) Photoreception in unicellular
flagellates: bioelectric phenomena in phototaxis, in Light in
Biology and Medicine (Douglas, R. D., ed) Plenum Press,
New York, pp. 523-532, doi: 10.1007/978-1-4684-5991-
3_53.
79. Sineshchekov, O. A. (1988) Phototaxis in microorganisms
and its role in photosynthesis regulation, in Phototrophic Mi-
croorganisms, Acad. Sci. USSR, Puschino, pp.11-18.
80. Yoshimura, K. (1994) Chromophore orientation in the
photoreceptor of Chlamydomonas as probed by stimulation
with polarized light, Photochem. Photobiol., 60, 594-597,
doi:10.1111/j.1751-1097.1994.tb05154.x.
81. Mast, S. O. (1911) Light and the Behavior of Organisms,
Wiley, New York, doi:10.5962/bhl.title.4601.
82. Morel-Laurens, N. M. L., and Feinleib, M. E. (1983)
Photomovement in an “eyeless” mutant of Chlamydo-
monas, Photochem. Photobiol., 37, 189-194, doi:10.1111/
j.1751-1097.1983.tb04457.x.
83. Kreimer, G., and Melkonian, M. (1990) Reflection con-
focal laser scanning microscopy of eyespot in flagellated
green algae, Eur. J. Cell Biol., 53, 101-111, doi:10.1016/
S1434-4610(99)70032-5.
84. Berthold, P., Tsunoda, S. P., Ernst, O. P., Mages, W.,
Gradmann, D., and Hegemann, P. (2008) Channel-
rhodopsin-1 initiates phototaxis and photophobic re-
sponses in Chlamydomonas by immediate light-induced
depolarization, Plant Cell, 20, 1665-1677, doi: 10.1105/
tpc.108.057919.
85. Mittelmeier, T. M., Thompson, M. D., Ozturk, E., and
Dieckmann, C. L. (2013) Independent localization of plas-
ma membrane and chloroplast components during eyespot
assembly, Eukaryot. Cell, 12, 1258-1270, doi: 10.1128/
EC.00111-13.
86. Schmidt, M., Gessner, G., Luff, M., Heiland, I., Wag-
ner,V., Kaminski, M., Geimer, S., Eitzinger, N., Reissen-
weber, T., Voytsekh, O., Fiedler, M., Mittag, M., and
Kreimer, G. (2006) Proteomic analysis of the eyespot of
Chlamydomonas reinhardtii provides novel insights into its
components and tactic movements, Plant Cell, 18, 1908-
1930, doi:10.1105/tpc.106.041749.
87. Melkonian, M., and Robenek, H. (1980) Eyespot mem-
branes of Chlamydomonas reinhardtii: a freeze-frac-
ture study, J. Ultrastruct. Res., 72, 90-102, doi: 10.1016/
S0022-5320(80)90138-0.
88. Awasthi, M., Ranjan, P., Sharma, K., Veetil, S. K., and
Kateriya, S. (2016) The trafficking of bacterial type rho-
dopsins into the Chlamydomonas eyespot and flagel-
la is IFT mediated, Sci. Rep., 6, 34646, doi: 10.1038/
srep34646.
89. Awasthi, M., Ranjan, P., and Kateriya, S. (2018) Cyto-
plasmic extensions of the channelrhodopsins1 and2 in-
teract in Chlamydomonas reinhardtii, J. Appl. Biotechnol.
Bioeng., 5, 84-90, doi:10.15406/jabb.2018.05.00121.
90. Sineshchekov, O. A., and Litvin, F. F. (1988) The mecha-
nisms of phototaxis in microorganisms, in Molecular Mecha-
nisms of Biological Action of Optical Radiation [in Russian],
Nauka, Moscow, pp.412-427.
91. Witman, G. B. (1993) Chlamydomonas phototaxis, Trends
Cell Biol., 3, 403-408, doi:10.1016/0962-8924(93)90091-e.
92. Kateriya, S., Nagel, G., Bamberg, E., and Hegemann, P.
(2004) “Vision” in single-celled algae, News Physiol. Sci.,
19, 133-137, doi:10.1152/nips.01517.2004.
93. Kreimer, G. (2009) The green algal eyespot apparatus:
a primordial visual system and more? Curr. Genet., 55,
19-43, doi:10.1007/s00294-008-0224-8.
94. Böhm, M., and Kreimer, G. (2021) Orient in the world
with a single eye: The green algal eyespot and phototax-
is, in Prog. Bot. (Cánovas, F. M., Lüttge, U., Risueño,
M.-C., and Pretzsch, H., eds). pp.259-304, doi:10.1007/
124_2020_38.
95. Sineshchekov, O. A. (1991) Electrophysiology of photo-
movements in flagellated algae. in Biophysics of Photore-
ceptors and Photomovements in Microorganisms (Lenci, F.,
Ghetti, F., Colombetti, G., Haeder, D.-P., and Song,
P.-S., eds) Plenum Press, New York. pp. 191-202,
doi:10.1007/978-1-4684-5988-3_15.
96. Rüffer, U., and Nultsch, W. (1990) Flagella photore-
sponses of Chlamydomonas cells held on micropipettes:
I. Change in flagellar beat frequency, Cell Motil. Cytoskele-
ton, 15, 162-167, doi:10.1002/cm.970150305.
97. Rüffer, U., and Nultsch, W. (1991) Flagellar photore-
sponses of Chlamydomonas cells held on micropipettes:
II.Change in flagellar beat pattern, Cell Motil. Cytoskele-
ton
, 18, 269-278, doi:10.1002/cm.970180404.
98. Josef, K., Saranak, J., and Foster, K.-W. (2005) Ciliary
behavior of a negatively phototactic Chlamydomonas rein-
hardtii, Cell Motil. Cytoskeleton, 61, 97-111, doi:10.1002/
cm.20069.
99. Josef, K., Saranak, J., and Foster, K.-W. (2006) Linear
systems analysis of the ciliary steering behavior associat-
ed with negative-phototaxis in Chlamydomonas reinhard-
tii, Cell Motil. Cytoskeleton, 63, 758-777, doi: 10.1002/
cm.20158.
CHANNELRHODOPSINS 1567
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
100. Holland, E.-M., Harz, H., Uhl, R., and Hegemann, P.
(1997) Control of phobic behavioral responses by rhodop-
sin-induced photocurrents in Chlamydomonas, Biophys.J.,
73, 1395-1401, doi:10.1016/S0006-3495(97)78171-2.
101. Schmidt, J. A., and Eckert, R. (1976) Calcium couples fla-
gella reversal to photostimulation in Chlamydomonas rein-
hardtii, Nature, 262, 713-715, doi:10.1038/262713a0.
102. Nultsch, W., and Häder, D.-P. (1979) Photomovement
of motile microorganisms, Photochem. Photobiol., 29,
423-437, doi:10.1111/j.1751-1097.1979.tb07072.x.
103. Harz, H., Nonnengässer, C., and Hegemann, P. (1992)
Thephotoreceptor current of the green alga Chlamydomo-
nas, Phil. Trans. R. Soc. Lond. B, 338, 39-52, doi:10.1098/
rstb.1992.0127.
104. Beck, C., and Uhl, R. (1994) On the localization of volt-
age-sensitive calcium channels in the flagella of Chlam-
ydomonas reinhardtii, J. Cell Biol., 125, 1119-1125,
doi:10.1083/jcb.125.5.1119.
105. Fujiu, K., Nakayama, Y., Yanagisawa, A., Sokabe, M.,
and Yoshimura, K. (2009) Chlamydomonas CAV2 encodes
a voltage-dependent calcium channel required for the
flagellar waveform conversion, Curr. Biol., 19, 133-139,
doi:10.1016/j.cub.2008.11.068.
106. Sineshchekov, O. A., Litvin, F. F., and Keszthelyi, L.
(1990) Two components of photoreceptor potential of the
flagellated green alga Haematococcus pluvialis, Biophys.J.,
57, 33-39, doi:10.1016/S0006-3495(90)82504-2.
107. Sineshchekov, O. A., Govorunova, E. G., and Spudich,
J.L. (2009) Photosensory functions of channelrhodopsins
in native algal cells, Photochem. Photobiol., 85, 556-563,
doi:10.1111/j.1751-1097.2008.00524.x.
108. Hegemann, P., and Marwan, W. (1988) Single photons
are sufficient to trigger movement responses in Chlam-
ydomonas reinhardtii, Photochem. Photobiol., 48, 99-106,
doi:10.1111/j.1751-1097.1988.tb02793.x.
109. Verret, F., Wheeler, G., Taylor, A. R., Farnham, G., and
Brownlee, C. (2010) Calcium channels in photosynthetic
eukaryotes: implications for evolution of calcium-based
signalling, New Phytol., 187, 23-43, doi: 10.1111 /
j.1469-8137.2010.03271.x.
110. Wagner, V., Ullmann, K., Mollwo, A., Kaminski, M.,
Mittag, M., and Kreimer, G. (2008) The phosphopro-
teome of a Chlamydomonas reinhardtii eyespot fraction
includes key proteins of the light signaling pathway, Plant
Physiol., 146, 772-788, doi:10.1104/pp.107.109645.
111. Linden, L., and Kreimer, G. (1995) Calcium modu-
lates rapid protein phosphorilation/dephosphorilation
in isolated eyespot apparatuses of the green alga Sper-
matozopsis similis, Planta, 197, 343-351, doi: 10.1007/
BF00202656.
112. Schlicher, U., Linden, L., Calenberg, M., and
Kreimer, G. (1995) G proteins and Ca
2+
modulated
protein kinases of a plasma membrane enriched frac-
tion and isolated eyespot apparatuses of Spermatozop-
sis similis (Chlorophycea), Eur. J. Phycol., 30, 319-330,
doi:10.1080/09670269500651111.
113. Calenberg, M., Brohnsonn, U., Zedlacher, M., and
Kreimer, G. (1998) Light- and Ca
2+
-modulated het-
eromeric GTPases in the eyespot apparatus of a flag-
ellate green alga, Plant Cell, 10, 91-103, doi: 10.1105/
tpc.10.1.91.
114. Trippens, J., Greiner, A., Schellwat, J., Neukam, M.,
Rottmann, T., Lu, Y., Kateriya, S., Hegemann, P., and
Kreimer, G. (2012) Phototropin influence on eyespot
development and regulation of phototactic behavior in
Chlamydomonas reinhardtii, Plant Cell, 24, 4687-4702,
doi:10.1105/tpc.112.103523.
115. Holland, E.-M., Braun, F.-J., Nonnengässer, C.,
Harz,H., and Hegemann,P. (1996) Thenature of rho-
dopsin-triggered photocurrents in Chlamydomonas. I.Ki-
netics and influence of divalent ions, Biophys. J., 70,
924-931, doi:10.1016/S0006-3495(96)79635-2.
116. Nonnengässer, C., Holland, E.-M., Harz, H., and
Hegemann, P. (1996) The nature of rhodopsin-activat-
ed photocurrents in Chlamydomonas. II. Influence of
monovalent ions, Biophys.J., 70, 932-938, doi:10.1016/
S0006-3495(96)79636-4.
117. Ritter, E., Stehfest, K., Berndt, A., Hegemann, P., and
Bartl, F. J. (2008) Monitoring light-induced structur-
al changes of Channelrhodopsin-2 by UV-visible and
Fourier transform infrared spectroscopy, J. Biol. Chem.,
283, 35033-35041, doi: 10.1074/jbc.M806353200.
118. Bamann, C., Kirsch, T., Nagel, G., and Bamberg, E.
(2008) Spectral characteristics of the photocycle of chan-
nelrhodopsin-2 and its implication for channel function,
J.Mol. Biol., 375, 686-694, doi:10.1016/j.jmb.2007.10.072.
119. Baidukova, O., Oppermann, J., Kelterborn, S., Fernandez
Lahore, R. G., Schumacher, D., Evers, H., Kamrani,
Y.Y., and Hegemann, P. (2022) Gating and ion selectiv-
ity of channelrhodopsins are critical for photo-activated
orientation of Chlamydomonas as shown by in vivo point
mutation, Nat. Commun., 13, 7253, doi:10.1038/s41467-
022-35018-6.
120. Dent, R. M., Haglund, C. M., Chin, B. L., Kobayashi,
M. C., and Niyogi, K.K. (2005) Functional genomics of
eukaryotic photosynthesis using insertional mutagenesis of
Chlamydomonas reinhardtii, Plant Physiol., 137, 545-556,
doi:10.1104/pp.104.055244.
121. Sizova, I., Greiner, A., Awasthi, M., Kateriya, S., and
Hegemann, P. (2013) Nuclear gene targeting in Chlam-
ydomonas using engineered zinc-finger nucleases, PlantJ.,
73, 873-882, doi:10.1111/tpj.12066.
122. Greiner, A., Kelterborn, S., Evers, H., Kreimer, G.,
Sizova, I., and Hegemann, P. (2017) Targeting of photore-
ceptor genes in Chlamydomonas reinhardtii via zinc-finger
nucleases and CRISPR/Cas9, Plant Cell, 29, 2498-2518,
doi:10.1105/tpc.17.00659.
123. Sineshchekov, O. A., Spudich, E. N., Trivedi, V. D., and
Spudich, J.L. (2006) Role of the cytoplasmic domain in
Anabaena sensory rhodopsin photocycling: vectoriality
of Schiff base deprotonation, Biophys.J., 91, 4519-4527,
doi:10.1529/biophysj.106.093641.
GOVORUNOVA, SINESHCHEKOV1568
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
124. Tashiro, R., Sushmita, K., Hososhima, S., Sharma, S.,
Kateriya, S., Kandori, H., and Tsunoda, S. P. (2021)
Specific residues in the cytoplasmic domain modulate
photocurrent kinetics of channelrhodopsin from the alga
Klebsormidium nitens, Commun. Biol., 4, 235, doi:10.1038/
s42003-021-01755-5.
125. Böhm, M., Boness, D., Fantisch, E., Erhard, H., Frauen-
holz, J., Kowalzyk, Z., Marcinkowski, N., Kateriya, S.,
Hegemann, P., and Kreimer, G. (2019) Channelrhodop-
sin-1 phosphorylation changes with the phototactic behav-
ior and responds to physiological stimuli in Chlamydomo-
nas, Plant Cell, 31, 886-910, doi:10.1105/tpc.18.00936.
126. Awasthi, M., Sushmita, K., Kaushik, M. S., Ranjan, P.,
and Kateriya, S. (2020) Novel modular rhodopsins from
green algae hold great potential for cellular optogenetic
modulation across the biological model systems, Life, 10,
259, doi:10.3390/life10110259.
127. Rozenberg, A., Oppermann, J., Wietek, J., Fernandez
Lahore, R. G., Sandaa, R. A., Bratbak, G., Hegemann,P.,
and Béjà, O. (2020) Lateral gene transfer of anion-con-
ducting channelrhodopsins between green algae and gi-
ant viruses, Curr. Biol., 30, 4910-4920, doi: 10.1016/
j.cub.2020.09.056.
128. Sharma, K., Sizova, I., Sanyal, S. K., Pandey, G. K.,
Hegemann, P., and Kateriya, S. (2023) Deciphering the
role of cytoplasmic domain of Channelrhodopsin in mod-
ulating the interactome and SUMOylome of Chlamydo-
monas reinhardtii, Int. J. Biol. Macromol., 243, 125135,
doi:10.1016/j.ijbiomac.2023.125135.
129. Matasci, N., Hung, L. H., Yan, Z., Carpenter, E. J.,
Wickett, N. J., Mirarab, S., Nguyen, N., Warnow, T.,
Ayyampalayam, S., Barker, M., Burleigh, J. G., Gitzen-
danner, M. A., Wafula, E., Der, J. P., dePamphilis, C.
W., Roure,B., Philippe, H., Ruhfel, B. R., Miles, N. W.,
Graham, S. W., et al. (2014) Data access for the 1,000
Plants (1KP) project, Gigascience, 3, 17, doi: 10.1186/
2047-217X-3-17.
130. Wong, G. K., Soltis, D. E., Leebens-Mack, J., Wickett,
N. J., Barker, M. S., Van de Peer, Y., Graham, S. W.,
and Melkonian, M. (2020) Sequencing and analyzing the
transcriptomes of a thousand species across the tree of
life for green plants, Annu. Rev. Plant Biol., 71, 741-765,
doi:10.1146/annurev-arplant-042916-041040.
131. Govorunova, E. G., Sineshchekov, O. A., Li, H., Wang, Y.,
Brown, L.S., Palmateer, A., Melkonian, M., Cheng, S.,
Carpenter, E., Patterson, J., Wong, G. K. S., and Spu-
dich, J.L. (2021) Cation and anion channelrhodopsins:
Sequence motifs and taxonomic distribution, MBio, 12,
e0165621, doi:10.1128/mBio.01656-21.
132. Glock, C., Nagpal, J., and Gottschalk, A. (2015) Micro-
bial rhodopsin optogenetic tools: Application for anal-
yses of synaptic transmission and of neuronal network
activity in behavior, Methods Mol. Biol., 1327, 87-103,
doi:10.1007/978-1-4939-2842-2_8.
133. Tsunoda, S. P., and Hegemann, P. (2009) Glu87 of chan-
nelrhodopsin-1 causes pH-dependent color tuning and
fast photocurrent inactivation, Photochem. Photobiol., 85,
564-569, doi:10.1111/j.1751-1097.2008.00519.x.
134. Govorunova, E. G., Sineshchekov, O. A., and Spudich,
J.L. (2017) Three families of channelrhodopsins and their
use in optogenetics, Zhurn. Visshej Nervnoi Dejatel’nosti,
67, 1-9.
135. Watanabe, M., and Furuya, M. (1982) Phototactic be-
havior of individual cells of Cryptomonas sp. in response
to continuous and intermittent light stimuli, Photo-
chem. Photobiol., 35, 559-563, doi: 10.1111/j.1751-1097.
1982.tb02609.x.
136. Rhiel, E., Häder, D.-P., and Wehrmeyer, W. (1988)
Photo-orientation in a freshwater Cryptomonas species,
J.Photochem. Photobiol. B Biol., 2, 123-132, doi:10.1016/
1011-1344(88)85041-3.
137. Erata, M., Kubota, M., Takahashi, T., Inouye, I., and
Watanabe, M. (1995) Ultrastructure and phototactic ac-
tion spectra of two genera of cryptophyte flagellate algae,
Cryptomonas and Chroomonas, Protoplasma, 188, 258-266,
doi:10.1007/BF01280378.
138. Govorunova, E. G., Sineshchekov, O. A., and Spudich,
J. L. (2016) Structurally distinct cation channelrhodop-
sins from cryptophyte algae, Biophys.J., 110, 2302-2304,
doi:10.1016/j.bpj.2016.05.001.
139. Gradinaru, V., Zhang, F., Ramakrishnan, C., Mattis, J.,
Prakash, R., Diester, I., Goshen, I., Thompson, K. R.,
and Deisseroth, K. (2010) Molecular and cellular ap-
proaches for diversifying and extending optogenetics, Cell,
141, 154-165, doi:10.1016/j.cell.2010.02.037.
140. Yamauchi, Y., Konno, M., Ito, S., Tsunoda, S. P., In-
oue, K., and Kandori, H. (2017) Molecular properties of
a DTD channelrhodopsin from Guillardia theta, Biophys.
Physicobiol., 14, 57-66, doi:10.2142/biophysico.14.0_57.
141. Hososhima, S., Ueno, S., Okado, S., Inoue, K. I.,
Konno,M., Yamauchi, Y., Inoue, K., Terasaki, H., Kan-
dori, H., and Tsunoda, S. P. (2023) Alight-gated cation
channel with high reactivity to weak light, Sci Rep., 13,
7625, doi:10.1038/s41598-023-34687-7.
142. Sineshchekov, O. A., Govorunova, E. G., Li, H.,
Wang,Y., Melkonian, M., Wong, G. K.-S., Brown, L.S.,
and Spudich, J. L. (2020) Conductance mechanisms of
rapidly desensitizing cation channelrhodopsins from
cryptophyte algae, mBio, 11, e00657-20, doi: 10.1128/
mBio.00657-20.
143. Kishi, K. E., Kim, Y. S., Fukuda, M., Inoue, M.,
Kusakizako, T., Wang, P. Y., Ramakrishnan, C., Byrne,
E. F. X., Thadhani, E., Paggi, J. M., Matsui, T. E.,
Yamashita, K., Nagata, T., Konno, M., Quirin, S., Lo,M.,
Benster, T., Uemura, T., Liu, K., Shibata, M., et al. (2022)
Structural basis for channel conduction in the pump-like
channelrhodopsin ChRmine, Cell, 185, 672-689.e623,
doi:10.1016/j.cell.2022.01.007.
144. Tucker, K., Sridharan, S., Adesnik, H., and Brohawn,
S.G. (2022) Cryo-EM structures of the channelrhodop-
sin ChRmine in lipid nanodiscs, Nat. Commun., 13, 4842,
doi:10.1038/s41467-022-32441-7.
CHANNELRHODOPSINS 1569
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
145. Kato, H. E., Zhang, F., Yizhar, O., Ramakrishnan, C.,
Nishizawa, T., Hirata, K., Ito, J., Aita, Y., Tsukazaki, T.,
Hayashi, S., Hegemann, P., Maturana, A. D., Ishitani, R.,
Deisseroth, K., and Nureki, O. (2012) Crystal structure of
the channelrhodopsin light-gated cation channel, Nature,
482, 369-374, doi:10.1038/nature10870.
146. Volkov, O., Kovalev, K., Polovinkin, V., Borshchevskiy,V.,
Bamann, C., Astashkin, R., Marin, E., Popov, A., Balan-
din, T., Willbold, D., Buldt, G., Bamberg, E., and Gor-
deliy, V. (2017) Structural insights into ion conduction by
channelrhodopsin2, Science, 358, eaan8862, doi:10.1126/
science.aan8862.
147. Oda, K., Vierock, J., Oishi, S., Rodriguez-Rozada, S.,
Taniguchi, R., Yamashita, K., Wiegert, J. S., Nishiza-
wa, T., Hegemann, P., and Nureki, O. (2018) Crystal
structure of the red light-activated channelrhodopsin
Chrimson, Nat. Commun., 9, 3949, doi:10.1038/s41467-
018-06421-9.
148. Marshel, J. H., Kim, Y. S., Machado, T. A., Quirin,S.,
Benson, B., Kadmon, J., Raja, C., Chibukhchyan, A.,
Ramakrishnan, C., Inoue, M., Shane, J. C., McKnight,
D. J., Yoshizawa, S., Kato, H. E., Ganguli, S., and
Deisseroth, K. (2019) Cortical layer-specific critical dy-
namics triggering perception, Science, 365, eaaw5202,
doi:10.1126/science.aaw5202.
149. Chen, R., Gore, F., Nguyen, Q. A., Ramakrishnan, C.,
Patel, S., Kim, S. H., Raffiee, M., Kim, Y. S., Hsueh,B.,
Krook-Magnusson, E., Soltesz, I., and Deisseroth, K.
(2020) Deep brain optogenetics without intracranial sur-
gery, Nat. Biotechnol., 39, 161-164, doi: 10.1038/s41587-
020-0679-9.
150. Govorunova, E. G., Sineshchekov, O. A., Rodarte, E. M.,
Janz, R., Morelle, O., Melkonian, M., Wong, G. K.-S.,
and Spudich, J.L. (2017) The expanding family of natu-
ral anion channelrhodopsins reveals large variations in
kinetics, conductance, and spectral sensitivity, Sci. Rep.,
7, 43358, doi:10.1038/srep43358.
151. Govorunova, E. G., Sineshchekov, O. A., Hemmati,R.,
Janz, R., Morelle, O., Melkonian, M., Wong, G.K. S.,
and Spudich, J. L. (2018) Extending the time domain
of neuronal silencing with cryptophyte anion chan-
nelrhodopsins, eNeuro, 5, ENEURO.0174-0118.2018,
doi:10.1523/ENEURO.0174-18.2018.
152. Dolgikh, D. A., Malyshev, A. Y., Salozhin, S. V., Nekraso-
va, O. V., Petrovskaya, L. E., Roshchin, M. V., Borodino-
va, A. A., Feldman, T. B., Balaban, P. M., Kirpichnikov,
M. P., and Ostrovsky, M.A. (2015) Anion-selective chan-
nelrhodopsin expressed in neuronal cell culture and in vivo
in murine brain: light-induced inhibition of generation of
action potentials, Dokl. Biochem. Biophys., 465, 424-427,
doi: 10.1134/S160767291506023X.
153. Mohammad, F., Stewart, J. C., Ott, S., Chlebikova, K.,
Chua, J. Y., Koh, T. W., Ho, J., and Claridge-Chang, A.
(2017) Optogenetic inhibition of behavior with anion chan-
nelrhodopsins, Nat. Methods, 14, 271-274, doi: 10.1038/
nmeth.4148.
154. Mohamed, G. A., Cheng, R. K., Ho, J., Krishnan, S.,
Mohammad, F., Claridge-Chang, A., and Jesuthasan,S.
(2017) Optical inhibition of larval zebrafish behaviour
with anion channelrhodopsins, BMC Biol., 15, 103,
doi:10.1186/s12915-017-0430-2.
155. Bergs, A., Schultheis, C., Fischer, E., Tsunoda, S. P.,
Erbguth, K., Husson, S. J., Govorunova, E.G., Spudich,
J.L., Nagel, G., Gottschalk, A., and Liewald, J.F. (2018)
Rhodopsin optogenetic toolbox v2.0 for light-sensitive
excitation and inhibition in Caenorhabditis elegans, PLoS
One, 13, e0191802, doi:10.1371/journal.pone.0191802.
156. Wilson, D. E., Scholl, B., and Fitzpatrick, D. (2018)
Differential tuning of excitation and inhibition shapes
direction selectivity in ferret visual cortex, Nature, 560,
97-101, doi:10.1038/s41586-018-0354-1.
157. Forli, A., Vecchia, D., Binini, N., Succol, F., Bovetti, S.,
Moretti, C., Nespoli, F., Mahn, M., Baker, C. A., Bolton,
M. M., Yizhar, O., and Fellin, T. (2018) Two-photon bidi-
rectional control and imaging of neuronal excitability with
high spatial resolution invivo, Cell Rep., 22, 3087-3098,
doi:10.1016/j.celrep.2018.02.063.
158. Andrei, A. R., Debes, S., Chelaru, M., Liu, X., Ro-
darte, E., Spudich, J. L., Janz, R., and Dragoi, V.
(2021) Heterogeneous side-effects of cortical inactiva-
tion in behaving animals, Elife, 10, e66400, doi:10.7554/
eLife.66400.
159. Huang, S., Ding, M., Roelfsema, M. R. G., Dreyer, I.,
Scherzer, S., Al-Rasheid, K. A. S., Gao, S., Nagel, G.,
Hedrich, R., and Konrad, K.R. (2021) Optogenetic con-
trol of the guard cell membrane potential and stomatal
movement by the light-gated anion channel GtACR1, Sci.
Adv., 7, doi:10.1126/sciadv.abg4619.
160. Zhou, Y., Ding, M., Gao, S., Yu-Strzelczyk, J.,
Krischke,M., Duan, X., Leide, J., Riederer, M., Muel-
ler, M. J., Hedrich, R., Konrad, K. R., and Nagel, G.
(2021) Optogenetic control of plant growth by a microbial
rhodopsin, Nat. Plants, 7, 144-151, doi: 10.1038/s41477-
021-00853-w.
161. Oppermann, J., Fischer, P., Silapetere, A., Liepe, B., Ro-
driguez-Rozada, S., Flores-Uribe, J., Peter, E., Keidel,A.,
Vierock, J., Kaufmann, J., Broser, M., Luck,M., Bartl,F.,
Hildebrandt, P., Simon Wiegert, J., Beja, O., Hege-
mann,P., and Wietek, J. (2019) MerMAIDs: a family of
metagenomically discovered marine anion-conducting
and intensely desensitizing channelrhodopsins, Nat. Com-
mun., 10, 3315, doi:10.1038/s41467-019-11322-6.
162. Malyshev, A. Y., Roshchin, M. V., Smirnova, G. R., Dol-
gikh, D. A., Balaban, P. M., and Ostrovsky, M. A. (2017)
Chloride conducting light activated channel GtACR2 can
produce both cessation of firing and generation of action
potentials in cortical neurons in response to light, Neuros-
ci. Lett., 640, 76-80, doi:10.1016/j.neulet.2017.01.026.
163. Messier, J. E., Chen, H., Cai, Z. L., and Xue, M. (2018)
Targeting light-gated chloride channels to neuronal som-
atodendritic domain reduces their excitatory effect in the
axon, Elife, 7, e38506, doi:10.7554/eLife.38506.
GOVORUNOVA, SINESHCHEKOV1570
BIOCHEMISTRY (Moscow) Vol. 88 No. 10 2023
164. Mahn, M., Prigge, M., Ron, S., Levy, R., and Yizhar, O.
(2016) Biophysical constraints of optogenetic inhibition
at presynaptic terminals, Nat. Neurosci., 19, 554-556,
doi:10.1038/nn.4266.
165. Mahn, M., Gibor, L., Patil, P., Cohen-Kashi Malina,K.,
Oring, S., Printz, Y., Levy, R., Lampl, I., and Yizhar,O.
(2018) High-efficiency optogenetic silencing with so-
ma-targeted anion-conducting channelrhodopsins, Nat.
Commun., 9, 4125, doi:10.1038/s41467-018-06511-8.
166. Mardinly, A. R., Oldenburg, I. A., Pegard, N. C., Sri-
dharan, S., Lyall, E. H., Chesnov, K., Brohawn, S. G.,
Waller, L., and Adesnik, H. (2018) Precise multimodal
optical control of neural ensemble activity, Nat. Neurosci.,
21, 881-893, doi:10.1038/s41593-018-0139-8.
167. Vierock, J., Peter, E., Grimm, C., Rozenberg, A., Chen,
I. W., Tillert, L., Castro Scalise, A. G., Casini, M.,
Augustin, S., Tanese, D., Forget, B. C., Peyronnet, R.,
Schneider- Warme, F., Emiliani, V., Béjà, O., and He-
gemann, P. (2022) WiChR, a highly potassium selec-
tive channelrhodopsin for low-light one- and two-pho-
ton inhibition of excitable cells, Sci. Adv., 8, eadd7729,
doi:10.1126/sciadv.add7729.
168. Govorunova, E. G., Sineshchekov, O. A., Brown, L. S.,
Bondar, A. N., and Spudich, J. L. (2022) Structural foun-
dations of potassium selectivity in channelrhodopsins,
mBio, 13, e0303922, doi:10.1128/mbio.03039-22.
169. Fan, L. Z., Kim, D. K., Jennings, J. H., Tian, H., Wang,
P. Y., Ramakrishnan, C., Randles, S., Sun, Y., Thad-
hani, E., Kim, Y. S., Quirin, S., Giocomo, L., Cohen,
A. E., and Deisseroth, K. (2023) All-optical physiology
resolves a synaptic basis for behavioral timescale plasticity,
Cell, 186, 543-559, doi:10.1016/j.cell.2022.12.035.