Next Article in Journal
Nanosheet-Facilitated Spray Delivery of dsRNAs Represents a Potential Tool to Control Rhizoctonia solani Infection
Previous Article in Journal
Exploring Next Generation Probiotics for Metabolic and Microbiota Dysbiosis Linked to Xenobiotic Exposure: Holistic Approach
Previous Article in Special Issue
Optical Activation of TrkB (E281A) in Excitatory and Inhibitory Neurons of the Mouse Visual Cortex
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Editorial

Light-Controlled Modulation and Analysis of Neuronal Functions

by
Carlo Matera
1,* and
Piotr Bregestovski
2,3,4,*
1
Department of Pharmaceutical Sciences, University of Milan, 20133 Milan, Italy
2
Institut National de la Santé et de la Recherche Médicale, Institut de Neurosciences des Systèmes, Aix-Marseille University, 13005 Marseille, France
3
Institute of Neurosciences, Kazan State Medical University, 420111 Kazan, Russia
4
Department of Normal Physiology, Kazan State Medical University, 420111 Kazan, Russia
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(21), 12921; https://doi.org/10.3390/ijms232112921
Submission received: 20 October 2022 / Accepted: 21 October 2022 / Published: 26 October 2022
(This article belongs to the Special Issue Light-Controlled Modulation and Analysis of Neuronal Functions)
Light is an extraordinary tool allowing us to read out and control neuronal functions thanks to its unique properties: it has a great degree of bioorthogonality and is minimally invasive; it can be precisely delivered with high spatial and temporal precision; and it can be used simultaneously or consequently at multiple wavelengths and locations. In the last 15 years, light-based methods have revolutionized the way we analyse and control biological systems. It turns out that with the help of light it is possible to study the functions of cells and cell ensembles [1,2,3], regulate the functions of deoxyribonucleic acid (DNA) [4], perform the photocontrol of peptide conformations [5], modulate the activity of voltage-gated and receptor-operated channels [6,7,8], and measure the concentrations of ions [9,10] and other cellular components [11,12]. It has become possible to control the behaviour of organisms [13,14], as well as to explore new ways of treating certain diseases [15,16,17]. Three novel approaches have been emerging more than others: optogenetics, optosensorics, and photopharmacology (optopharmacology).
Optogenetics is an elegant method that combines optical technology and genetic engineering to control the functions of biological systems (cells, tissues, organs, organisms) genetically modified to express photosensitive proteins [2]. By stimulating with light, this method provides high-spatiotemporal and high-specificity resolutions, in contrast to conventional pharmacological or electrical stimulation [18,19]. Optogenetics provides a route to study synaptic circuits [20] and underlying movement diseases [21,22] and has become an effective technology to revolutionize brain research for the therapy of vision [23,24,25], cardiovascular [26], and neurodegenerative disorders [27,28]. It has also been extended to other biomedical fields [29,30].
Optosensorics is the direction of research exploring the possibility of non-invasively monitoring intracellular ion concentrations and the activity of enzymes, lipids, and other cellular components using specific optical sensors. Most of the sensors developed in the past years are genetically encoded tools. These probes possess fluorophore groups capable of changing fluorescence when interacting with certain ions or molecules [31]. The use of modern optical and fluorescent technologies in combination with molecular genetic techniques has provided the possibility for the biosensoric monitoring of reactive oxidative species [32,33], cAMP [34], glucose [35,36], glutamate [37,38], pyruvate [39], lactate [40], different ions [41,42], and neuronal activity [43].
In contrast to the above-mentioned methods, photopharmacology relies on synthetic photoswitchable ligands that are externally supplied and usually does not require genetic manipulation. Photopharmacological agents are obtained via the incorporation of a molecular photoswitch (e.g., azobenzene and its derivatives) into the structure of a biologically active compound, so that light can be used as an external signal to “switch” the ligand conformation and hence its activity between two different states (ideally “off” and “on”) [44,45]. This is ultimately performed to achieve control over biological systems with a high degree of temporal and spatial precision for investigational and therapeutic purposes. Potential applications of photopharmacology include neurobiology [46,47,48], cancer [49,50], microbial infections [51], pain [52,53], and blindness [54,55], among others.
In this Special Issue, several papers focus on neurobiological aspects, using light as a tool for experimental analysis. In the paper published by Gerasimov et al., the authors describe an optogenetic approach to stimulate astrocytes with the aim to modulate neuronal activity [56]. They applied light stimulation to astrocytes expressing a version of ChR2 (ionotropic opsin) or Opto-α1AR (metabotropic opsin). Optimal optogenetic stimulation parameters were determined using patch-clamp recordings of hippocampal pyramidal neurons’ spontaneous activity in brain slices as a readout. They observed that the activation of the astrocytic Opto-α1AR, but not ChR2, results in an increase in the fEPSP slope in hippocampal neurons. The authors conclude that Opto-α1AR expressed in hippocampal astrocytes provides an opportunity to modulate the long-term synaptic plasticity optogenetically and suggest that this approach may potentially be used to normalize the synaptic transmission and plasticity defects in a variety of neuropathological conditions, including models of Alzheimer’s disease and other neurodegenerative disorders.
In another work, Abd El-Aziz et al. used an optogenetic system to study the activity of the epithelial Na+ channel (ENaC) [57]. The activity of ENaC is strongly dependent on the membrane phospholipid phosphatidylinositol 4,5-bisphosphate (PIP2). PIP2 binds two distinct cationic clusters within the N termini of β- and γ-ENaC subunits (βN1 and γN2). The purpose of this study was to determine whether each independent PIP2–ENaC interaction site is sufficient to abolish the response of ENaC to changes in PIP2 levels. The authors had previously determined the affinities of these sites using short synthetic peptides. In this paper, they describe their role in sensitizing ENaC to changes in PIP2 levels in the cellular system. For this purpose, they compared the effects of PIP2 depletion and recovery on ENaC channel activity and intracellular Na+ levels [Na+]i. They tested the effects on ENaC activity with mutations to the PIP2 binding sites using the optogenetic system CIBN/CRY2-OCRL to selectively deplete PIP2. Whole cell patch-clamp measurements showed a complete lack of response to PIP2 depletion and recovery in ENaC with mutations to βN1 or γN2 or both sites compared to wild-type ENaC. These results suggest that the βN1 and γN2 sites on ENaC are each necessary to permit maximal ENaC activity in the presence of PIP2.
Lilja et al. presented the results obtained with a novel optogenetic tool for the investigation of neuroplasticity [58]. The activation of tropomyosin receptor kinase B (TrkB), the receptor of brain-derived neurotrophic factor, plays a key role in induced juvenile-like plasticity (iPlasticity), which allows the restructuring of neural networks in adulthood. In this work, they evaluated the utility of a new, highly sensitive, optically activatable tropomyosin receptor kinase B receptor, named OptoTrkB (E281A). OptoTrkB (E281A) was successfully transduced in parvalbumin-positive (PV+) interneurons and in alpha calcium/calmodulin-dependent protein kinase type II positive (CKII+) pyramidal neurons specifically. Light stimulation through transparent skulls or even through a high-opacity barrier (intact skull and fur) promoted the phosphorylation of ERK and CREB, downstream signals of TrkB, in the neurons expressing optoTrkB (E281A) at a certain level. Their results indicate that this highly sensitive optoTrkB (E281A) receptor can be activated using wireless optogenetic methods and can thus be used for a broad range of behavioural studies. Overall, these findings show that the highly sensitive optoTrkB (E281A) can be used in iPlasticity studies of both inhibitory and excitatory neurons, with flexible stimulation protocols in behavioural studies.
The application of novel optogenetic/chemogenetic tools and advanced in vitro models, including those based on iPSC-derived cells, organoids, or utilizing 3D brain-on-chip platforms, are of great importance for the development of new therapeutic options and the assessment of aberrant neurogenesis in Parkinson’s-type neurodegeneration. In their review article, Salmina et al. discuss current approaches to assess neurogenesis and prospects in the application of optogenetic protocols to restore neurogenesis in patients with Parkinson’s disease [59].
In their paper, Ivashkina et al. present a methodological development and experimental application of a genetically engineered optosensoric tool that allows the imaging of activity-induced neuronal c-fos expression [60]. The authors show a novel light-controlled approach for the long-term analysis of calcium activity in the cortical neurons that were specifically tagged through c-fos expression during a particular cognitive episode. In addition, using in vivo two-photon imaging of Fos-GFP PtA neurons, they report specific changes in neuronal activity in this cortical area during both the acquisition and retrieval of associative fear memory. Taken together, these results suggest that Fos-Cre-GCaMP mice are suitable for the investigation of calcium activity in the neurons which were specifically activated during a particular learning episode.
In the work by Zhilyakov et al., the authors reported the results from their investigation on the relationship between the nicotine-induced autoregulation of acetylcholine (ACh) release and the changes in the concentration of presynaptic calcium levels [61]. For this purpose, they used a pharmacological approach, electrophysiological techniques, and a method for the optical registration of changes in calcium levels in the motor nerve endings. The authors found that an agonist of nicotinic receptors (at a concentration not significantly affecting the state of the postsynaptic membrane) leads to a decrease in the amount of released ACh quanta. This effect is accompanied not by a decrease but by an increase in calcium ion entry into the motor nerve terminal. These results suggest that the nicotinic cholinergic receptors responsible for the mechanism of ACh release autoregulation are the receptors of neuronal type. Activation of these receptors leads to the upregulation of the Cav1 type of VGCCs, resulting in the enhancement of Ca2+ entry into the nerve ending. This study adds new elements to the understanding of the cholinergic system functioning and the envisioning of novel potential approaches for the treatment of diseases associated with cholinergic dysfunctions.
In the field of optosensorics, Ponomareva et al. described the effectiveness of the genetically encoded biosensor, ClopHensor, expressed in transgenic mice for the estimation of [H+]i and [Cl]i concentrations in brain slices [62]. They performed simultaneous monitoring of [H+]i and [Cl]i under different experimental conditions, including changing external concentrations of ions (Ca2+, Cl, K+, Na+) and the synaptic stimulation of Shaffer’s collaterals of hippocampal slices. The results obtained illuminate the different pathways regulating Cl and pH equilibrium in neurons and demonstrate that ClopHensor, expressed in transgenic mice, represents an efficient tool for the non-invasive monitoring of intracellular Cl and H+ ions.
In their work, Sotskov et al. performed in vivo imaging of neuronal activity in the CA1 field of the mouse hippocampus using genetically encoded green calcium indicators, including the novel NCaMP7 and FGCaMP7, designed specifically for in vivo calcium imaging [63]. The purpose of this study was to investigate the dynamics of the initial place field formation in the mouse hippocampus. Their data show that neuronal activity recorded with genetically encoded calcium sensors revealed fast behaviour-dependent plasticity in the mouse hippocampus, resulting in the rapid formation of place fields and population activity that allowed the reconstruction of the geometry of the navigated maze. Taken together, these results reveal the fast emergence and tuning dynamics of place cell codes and demonstrate the applicability of novel calcium indicators NCaMP7 and FGCaMP7 in the light-controlled analysis of neural functions in behaving mice.
In the field of photopharmacology, Matera et al. reported a novel photoswitchable ligand that enables reversible spatiotemporal control of dopaminergic transmission [64]. They demonstrated that this new photoswitch, named azodopa, activates D1-like receptors in vitro in a light-dependent manner. Moreover, azodopa enables reversibly photocontrolling zebrafish motility on a timescale of seconds and allows separating the retinal component of dopaminergic neurotransmission. Finally, they proved that azodopa increases the overall neural activity in the cortex of anesthetized mice and displays illumination-dependent activity in individual neurons. Azodopa is the first photoswitchable dopamine agonist with demonstrated efficacy in wild-type animals and opens the way forward to remotely controlling dopaminergic neurotransmission for investigational and therapeutic purposes.
Finally, the review article by Nin-Hill et al. discusses the application of structure-based computational methods, such as homology modelling, molecular docking, molecular dynamics, and enhanced sampling techniques, to photoswitchable ligands targeting voltage- and ligand-gated ion channels [65]. Notably, the examples presented by the authors show how the integration of computational modelling with experimental data can greatly facilitate photoswitchable ligand design and optimization and provide structural insights to understand the observed light-regulated effects. They conclude by stating that the latest advances in structural biology will further support computer-assisted approaches in photopharmacology.
Overall, this Special Issue of the International Journal of Molecular Sciences contains original contributions reporting recent advances in the fields of optogenetics, optosensorics, and photopharmacology, as well as review papers discussing key achievements and prospects in the field. Therefore, we expect that these articles will be of interest to many scientists working with light-based biological methods and will inspire further investigations in relevant research areas.

Author Contributions

Conceptualization, C.M. and P.B.; writing—original draft preparation, C.M. and P.B.; writing—review and editing, C.M. and P.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Scientific Foundation (project no. 18-15-00313 to P.B.).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Deisseroth, K. Optogenetics. Nat. Methods 2011, 8, 26–29. [Google Scholar] [CrossRef] [PubMed]
  2. Deisseroth, K. Optogenetics: 10 years of microbial opsins in neuroscience. Nat. Neurosci. 2015, 18, 1213–1225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Kim, C.K.; Adhikari, A.; Deisseroth, K. Integration of optogenetics with complementary methodologies in systems neuroscience. Nat. Rev. Neurosci. 2017, 18, 222–235. [Google Scholar] [CrossRef] [PubMed]
  4. Kashida, H.; Liang, X.; Asanuma, H. Rational design of functional DNA with a non-ribose acyclic scaffold. Curr. Organ. Chem. 2009, 13, 1065–1084. [Google Scholar] [CrossRef]
  5. Szymański, W.; Beierle, J.M.; Kistemaker, H.A.; VelemNa, W.A.; Feringa, B.L. Reversible photocontrol of biological systems by the incorporation of molecular photoswitches. Chem. Rev. 2013, 113, 6114–6178. [Google Scholar] [CrossRef] [Green Version]
  6. Gorostiza, P.; Isacoff, E.Y. Optical switches for remote and noninvasive control of cell signaling. Science 2008, 322, 395–399. [Google Scholar] [CrossRef]
  7. Banghart, M.R.; Mourot, A.; Fortin, D.L.; Yao, J.Z.; Kramer, R.H.; Trauner, D. Photochromic blockers of voltage-gated potassium channels. Angew. Chem. Int. Ed. Engl. 2009, 48, 9097–9101. [Google Scholar] [CrossRef] [Green Version]
  8. Bregestovski, P.; Maleeva, G.; Gorostiza, P. Light-induced regulation of ligand-gated channel activity. Br. J. Pharmacol. 2018, 175, 1892–1902. [Google Scholar] [CrossRef] [Green Version]
  9. Bregestovski, P.; Waseem, T.; Mukhtarov, M. Genetically encoded optical sensors for monitoring of intracellular chloride and chloride-selective channel activity. Front. Mol. Neurosci. 2009, 2, 15. [Google Scholar] [CrossRef] [Green Version]
  10. Suzuki, J.; Kanemaru, K.; Iino, M. Genetically encoded fluorescent indicators for organellar calcium imaging. Biophys. J. 2016, 111, 1119–1131. [Google Scholar] [CrossRef]
  11. Imamura, H.; Nhat, K.P.H.; Togawa, H.; Saito, K.; Iino, R.; Kato-Yamada, Y.; Nagai, T.; Noji, H. Visualization of ATP levels inside single living cells with fluorescence resonance energy transfer-based genetically encoded indicators. Proc. Natl. Acad. Sci. USA 2009, 106, 15651–15656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Berg, J.; Hung, Y.P.; Yellen, G. A genetically encoded fluorescent reporter of ATP: ADP ratio. Nat. Methods 2009, 6, 161–166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Miesenböck, G. Optogenetic control of cells and circuits. Annu. Rev. Cell Dev. Biol. 2011, 27, 731–758. [Google Scholar] [CrossRef] [Green Version]
  14. Haubensak, W.; Kunwar, P.S.; Cai, H.; Ciocchi, S.; Wall, N.R.; Ponnusamy, R.; Biag, J.; Dong, H.W.; Deisseroth, K.; Callaway, E.M.; et al. Genetic dissection of an amygdala microcircuit that gates conditioned fear. Nature 2010, 468, 270–276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Rossi, M.A.; Calakos, N.; Yin, H.H. Spotlight on movement disorders: What optogenetics has to offer. Mov. Disord. 2015, 30, 624–631. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Häusser, M. Optogenetics—The Might of Light. N. Engl. J. Med. 2021, 385, 1623–1626. [Google Scholar] [CrossRef]
  17. Berry, M.H.; Holt, A.; Broichhagen, J.; Donthamsetti, P.; Flannery, J.G.; Isacoff, E.Y. Photopharmacology for vision restoration. Curr. Opin. Pharmacol. 2022, 65, 102259. [Google Scholar] [CrossRef]
  18. Tye, K.M.; Deisseroth, K. Optogenetic investigation of neural circuits underlying brain disease in animal models. Nat. Rev. Neurosci. 2012, 13, 251–266. [Google Scholar] [CrossRef]
  19. Häusser, M. Optogenetics: The age of light. Nat. Methods 2014, 11, 1012–1014. [Google Scholar] [CrossRef]
  20. Linders, L.E.; Supiot, L.; Du, W.; D’Angelo, R.; Adan, R.A.; Riga, D.; Meye, F.J. Studying Synaptic Connectivity and Strength with Optogenetics and Patch-Clamp Electrophysiology. Int. J. Mol. Sci. 2022, 23, 11612. [Google Scholar] [CrossRef]
  21. Fougère, M.; van der Zouwen, C.I.; Boutin, J.; Neszvecsko, K.; Sarret, P.; Ryczko, D. Optogenetic stimulation of glutamatergic neurons in the cuneiform nucleus controls locomotion in a mouse model of Parkinson’s disease. Proc. Natl. Acad. Sci. USA 2021, 118, e2110934118. [Google Scholar] [CrossRef] [PubMed]
  22. Parker, K.L.; Kim, Y.; Alberico, S.L.; Emmons, E.B.; Narayanan, N.S. Optogenetic approaches to evaluate striatal function in animal models of Parkinson disease. Dialogues Clin. Neurosci. 2016, 18, 99–107. [Google Scholar] [CrossRef] [PubMed]
  23. Chaffiol, A.; Duebel, J. Mini-review: Cell type-specific optogenetic vision restoration approaches. Adv. Exp. Med. Biol. 2018, 1074, 69–73. [Google Scholar] [PubMed]
  24. McClements, M.E.; Staurenghi, F.; MacLaren, R.E.; Cehajic-Kapetanovic, J. Optogenetic gene therapy for the degenerate retina: Recent advances. Front. Neurosci. 2020, 14, 570909. [Google Scholar] [CrossRef]
  25. 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.; et al. Partial recovery of visual function in a blind patient after optogenetic therapy. Nat. Med. 2021, 27, 1223–1229. [Google Scholar] [CrossRef]
  26. Nussinovitch, U.; Gepstein, L. Optogenetics for in vivo cardiac pacing and resynchronization therapies. Nat. Biotechnol. 2015, 33, 750–754. [Google Scholar] [CrossRef]
  27. Kastanenka, K.V.; Calvo-Rodriguez, M.; Hou, S.S.; Zhou, H.; Takeda, S.; Arbel-Ornath, M.; Lariviere, A.; Lee, Y.F.; Kim, A.; Hawkes, J.M.; et al. Frequency-dependent exacerbation of Alzheimer’s disease neuropathophysiology. Sci. Rep. 2019, 9, 8964. [Google Scholar] [CrossRef] [Green Version]
  28. Ko, H.; Yoon, S.P. Optogenetic neuromodulation with gamma oscillation as a new strategy for Alzheimer disease: A narrative review. J. Yeungnam Med. Sci. 2022, 39, 269–277. [Google Scholar] [CrossRef]
  29. Ye, H.; Fussenegger, M. Optogenetic medicine: Synthetic therapeutic solutions precision-guided by light. Cold Spring Harb. Perspect. Med. 2019, 9, a034371. [Google Scholar] [CrossRef] [Green Version]
  30. Patrono, E.; Svoboda, J.; Stuchlík, A. Schizophrenia, the gut microbiota, and new opportunities from optogenetic manipulations of the gut-brain axis. Behav. Brain Funct. 2021, 17, 7. [Google Scholar] [CrossRef]
  31. Bolbat, A.; Schultz, C. Recent developments of genetically encoded optical sensors for cell biology. Biology of the Cell 2017, 109, 1–23. [Google Scholar] [CrossRef] [PubMed]
  32. Belousov, V.V.; Fradkov, A.F.; Lukyanov, K.A.; Staroverov, D.B.; Shakhbazov, K.S.; Terskikh, A.V.; Lukyanov, S. Genetically encoded fluorescent indicator for intracellular hydrogen peroxide. Nat. Methods 2006, 3, 281–286. [Google Scholar] [CrossRef]
  33. Ermakova, Y.G.; Bilan, D.S.; Matlashov, M.E.; Mishina, N.M.; Markvicheva, K.N.; Subach, O.M.; Subach, F.V.; Bogeski, I.; Hoth, M.; Enikolopov, G.; et al. Red fluorescent genetically encoded indicator for intracellular hydrogen peroxide. Nat. Commun. 2014, 5, 5222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Ponsioen, B.; Zhao, J.; Riedl, J.; Zwartkruis, F.; van der Krogt, G.; Zaccolo, M.; Moolenaar, W.H.; Bos, J.L.; Jalink, K. Detecting cAMP-induced Epac activation by fluorescence resonance energy transfer: Epac as a novel cAMP indicator. EMBO Rep. 2004, 5, 1176–1180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Deuschle, K.; Okumoto, S.; Fehr, M.; Looger, L.L.; Kozhukh, L.; Frommer, W.B. Construction and optimization of a family of genetically encoded metabolite sensors by semirational protein engineering. Protein Sci. 2005, 14, 2304–2314. [Google Scholar] [CrossRef] [PubMed]
  36. Takanaga, H.; Chaudhuri, B.; Frommer, W.B. GLUT1 and GLUT9 as major contributors to glucose influx in HepG2 cells identified by a high sensitivity intramolecular FRET glucose sensor. Biochim. Biophys. Acta BBA Biomembr. 2008, 1778, 1091–1099. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Okumoto, S.; Looger, L.L.; Micheva, K.D.; Reimer, R.J.; Smith, S.J.; Frommer, W.B. Detection of glutamate release from neurons by genetically encoded surface-displayed FRET nanosensors. Proc. Natl. Acad. Sci. USA 2005, 102, 8740–8745. [Google Scholar] [CrossRef] [Green Version]
  38. Hires, S.A.; Zhu, Y.; Tsien, R.Y. Optical measurement of synaptic glutamate spillover and reuptake by linker optimized glutamate-sensitive fluorescent reporters. Proc. Natl. Acad. Sci. USA 2008, 105, 4411–4416. [Google Scholar] [CrossRef] [Green Version]
  39. San Martin, A.; Ceballo, S.; Baeza-Lehnert, F.; Lerchundi, R.; Valdebenito, R.; Contreras-Baeza, Y.; Alegria, K.; Barros, L.F. Imaging mitochondrial flux in single cells with a FRET sensor for pyruvate. PLoS ONE 2014, 9, e85780. [Google Scholar] [CrossRef] [Green Version]
  40. San Martín, A.; Ceballo, S.; Ruminot, I.; Lerchundi, R.; Frommer, W.B.; Barros, L.F. A genetically encoded FRET lactate sensor and its use to detect the Warburg effect in single cancer cells. PLoS ONE 2013, 8, e57712. [Google Scholar] [CrossRef]
  41. Tian, L.; Hires, S.A.; Mao, T.; Huber, D.; Chiappe, M.E.; Chalasani, S.H.; Petreanu, L.; Akerboom, J.; McKinney, S.A.; Schreiter, E.R.; et al. Imaging neural activity in worms, flies and mice with improved GCaMP calcium indicators. Nat. Methods 2019, 6, 875–881. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Markova, O.; Mukhtarov, M.; Real, E.; Jacob, Y.; Bregestovski, P. Genetically encoded chloride indicator with improved sensitivity. J. Neurosci. Methods 2008, 170, 67–76. [Google Scholar] [CrossRef] [PubMed]
  43. Shen, Y.; Nasu, Y.; Shkolnikov, I.; Kim, A.; Campbell, R.E. Engineering genetically encoded fluorescent indicators for imaging of neuronal activity: Progress and prospects. Neurosci. Res. 2002, 152, 3–14. [Google Scholar] [CrossRef] [PubMed]
  44. McKenzie, C.K.; Sanchez-Romero, I.; Janovjak, H. Flipping the Photoswitch: Ion Channels Under Light Control. Adv. Exp. Med. Biol. 2015, 869, 101–117. [Google Scholar]
  45. Broichhagen, J.; Frank, J.A.; Trauner, D. A roadmap to success in photopharmacology. Acc. Chem. Res. 2015, 48, 1947–1960. [Google Scholar] [CrossRef]
  46. Hull, K.; Morstein, J.; Trauner, D. In vivo photopharmacology. Chem. Rev. 2018, 118, 10710–10747. [Google Scholar] [CrossRef]
  47. Bregestovski, P.D.; Maleeva, G.V. Photopharmacology: A brief review using the control of potassium channels as an example. Neurosci. Behav. Physiol. 2019, 49, 184–191. [Google Scholar] [CrossRef]
  48. Castagna, R.; Kolarski, D.; Durand-de Cuttoli, R.; Maleeva, G. Orthogonal Control of Neuronal Circuits and Behavior Using Photopharmacology. J. Mol. Neurosci. 2022, 72, 1433–1442. [Google Scholar] [CrossRef]
  49. Szymanski, W.; Ourailidou, M.E.; Velema, W.A.; Dekker, F.J.; Feringa, B.L. Light-controlled histone deacetylase (HDAC) inhibitors: Towards photopharmacological chemotherapy. Chem. Eur. J. 2015, 21, 16517–16524. [Google Scholar] [CrossRef] [Green Version]
  50. Udasin, R.; Sil, A.; Zomot, E.; Achildiev Cohen, H.; Haj, J.; Engelmayer, N.; Lev, S.; Binshtok, A.M.; Shaked, Y.; Kienzler, M.A.; et al. Photopharmacological modulation of native CRAC channels using azoboronate photoswitches. Proc. Natl. Acad. Sci. USA 2022, 119, e2118160119. [Google Scholar] [CrossRef]
  51. Bispo, M.; van Dijl, J.M.; Szymanski, W. Molecular Photoswitches in Antimicrobial Photopharmacology. Mol. Photoswitches Chem. Prop. Appl. 2022, 2, 843–871. [Google Scholar]
  52. Gazerani, P. Shedding light on photo-switchable analgesics for pain. Pain Manag. 2017, 7, 71–74. [Google Scholar] [CrossRef] [PubMed]
  53. Iseppon, F.; Arcangeletti, M. Optogenetics and photopharmacology in pain research and therapeutics. STEMedicine 2020, 1, e43. [Google Scholar] [CrossRef]
  54. Caporale, N.; Kolstad, K.D.; Lee, T.; Tochitsky, I.; Dalkara, D.; Trauner, D.; Kramer, R.; Dan, Y.; Isacoff, E.Y.; Flannery, J.G. LiGluR restores visual responses in rodent models of inherited blindness. Mol. Ther. 2011, 19, 1212–1219. [Google Scholar] [CrossRef] [PubMed]
  55. Prischich, D.; Gomila, A.M.J.; Milla-Navarro, S.; Sangüesa, G.; Diez-Alarcia, R.; Preda, B.; Matera, C.; Batlle, M.; Ramírez, L.; Giralt, E.; et al. Adrenergic Modulation With Photochromic Ligands. Angew. Chem. Int. Ed. Engl. 2021, 60, 3625–3631. [Google Scholar] [CrossRef]
  56. Gerasimov, E.; Erofeev, A.; Borodinova, A.; Bolshakova, A.; Balaban, P.; Bezprozvanny, I.; Vlasova, O.L. Optogenetic Activation of Astrocytes-Effects on Neuronal Network Function. Int. J. Mol. Sci. 2021, 22, 9613. [Google Scholar] [CrossRef] [PubMed]
  57. Abd El-Aziz, T.M.; Kaur, A.; Shapiro, M.S.; Stockand, J.D.; Archer, C.R. Optogenetic Control of PIP2 Interactions Shaping ENaC Activity. Int. J. Mol. Sci. 2022, 23, 3884. [Google Scholar] [CrossRef]
  58. Lilja, A.; Didio, G.; Hong, J.; Heo, W.D.; Castrén, E.; Umemori, J. Optical Activation of TrkB (E281A) in Excitatory and Inhibitory Neurons of the Mouse Visual Cortex. Int. J. Mol. Sci. 2022, 23, 10249. [Google Scholar] [CrossRef]
  59. Salmina, A.B.; Kapkaeva, M.R.; Vetchinova, A.S.; Illarioshkin, S.N. Novel Approaches Used to Examine and Control Neurogenesis in Parkinson’s Disease. Int. J. Mol. Sci. 2021, 22, 9608. [Google Scholar] [CrossRef]
  60. Ivashkina, O.I.; Gruzdeva, A.M.; Roshchina, M.A.; Toropova, K.A.; Anokhin, K.V. Imaging of C-fos Activity in Neurons of the Mouse Parietal Association Cortex during Acquisition and Retrieval of Associative Fear Memory. Int. J. Mol. Sci. 2021, 22, 8244. [Google Scholar] [CrossRef]
  61. Zhilyakov, N.; Arkhipov, A.; Malomouzh, A.; Samigullin, D. Activation of Neuronal Nicotinic Receptors Inhibits Acetylcholine Release in the Neuromuscular Junction by Increasing Ca2+ Flux through Cav1 Channels. Int. J. Mol. Sci. 2021, 22, 9031. [Google Scholar] [CrossRef] [PubMed]
  62. Ponomareva, D.; Petukhova, E.; Bregestovski, P. Simultaneous Monitoring of pH and Chloride (Cl-) in Brain Slices of Transgenic Mice. Int. J. Mol. Sci. 2021, 22, 13601. [Google Scholar] [CrossRef] [PubMed]
  63. Sotskov, V.P.; Pospelov, N.A.; Plusnin, V.V.; Anokhin, K.V. Calcium Imaging Reveals Fast Tuning Dynamics of Hippocampal Place Cells and CA1 Population Activity during Free Exploration Task in Mice. Int. J. Mol. Sci. 2022, 23, 638. [Google Scholar] [CrossRef] [PubMed]
  64. Matera, C.; Calvé, P.; Casadó-Anguera, V.; Sortino, R.; Gomila, A.M.J.; Moreno, E.; Gener, T.; Delgado-Sallent, C.; Nebot, P.; Costazza, D.; et al. Reversible Photocontrol of Dopaminergic Transmission in Wild-Type Animals. Int. J. Mol. Sci. 2022, 23, 10114. [Google Scholar] [CrossRef]
  65. Nin-Hill, A.; Mueller, N.P.F.; Molteni, C.; Rovira, C.; Alfonso-Prieto, M. Photopharmacology of Ion Channels through the Light of the Computational Microscope. Int. J. Mol. Sci. 2021, 22, 12072. [Google Scholar] [CrossRef]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Matera, C.; Bregestovski, P. Light-Controlled Modulation and Analysis of Neuronal Functions. Int. J. Mol. Sci. 2022, 23, 12921. https://doi.org/10.3390/ijms232112921

AMA Style

Matera C, Bregestovski P. Light-Controlled Modulation and Analysis of Neuronal Functions. International Journal of Molecular Sciences. 2022; 23(21):12921. https://doi.org/10.3390/ijms232112921

Chicago/Turabian Style

Matera, Carlo, and Piotr Bregestovski. 2022. "Light-Controlled Modulation and Analysis of Neuronal Functions" International Journal of Molecular Sciences 23, no. 21: 12921. https://doi.org/10.3390/ijms232112921

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop