The islets of Langerhans reside within the endocrine pancreas as highly vascularized microorgans that are responsible for the secretion of key hormones, such as insulin and glucagon. Islet function relies on a range of dynamic molecular processes that include Ca2+ waves, hormone pulses, and complex interactions between islet cell types. Dysfunction of these processes results in poor maintenance of blood glucose homeostasis and is a hallmark of diabetes. Recently, the development of optogenetic methods that rely on light-sensitive molecular actuators has allowed perturbation of islet function with near physiological spatiotemporal acuity. These actuators harness natural photoreceptor proteins and their engineered variants to manipulate mouse and human cells that are not normally light-responsive. Until recently, optogenetics in islet biology has primarily focused on controlling hormone production and secretion; however, studies on further aspects of islet function, including paracrine regulation between islet cell types and dynamics within intracellular signaling pathways, are emerging. Here, we discuss the applicability of optogenetics to islets cells and comprehensively review seminal as well as recent work on optogenetic actuators and their effects in islet function and diabetes mellitus.

Article Highlights
  • Optogenetics harnesses light-sensitive actuators to perturb biological systems with unparalleled spatiotemporal precision.

  • Here, we comprehensively review recent and seminal applications of optogenetics to islets and designer islet-like cells.

  • We cover hormone production and secretion, cell-cell communication, and intracellular signaling dynamics, as well as discuss the merits and limitations of the technique.

  • Future opportunities focus on a broad applicability of the method and potential clinical perspectives.

The field of optogenetics integrates genetic and optical principles to manipulate cell and animal physiology using light. Through genetic targeting, light-sensitive photoreceptor proteins are expressed in defined cell types so that specific behaviors can be remote controlled with unprecedented spatiotemporal precision. Optogenetics originally emerged as a method to manipulate action potential firing in neural circuits (1,2) but has been increasingly applied in other contexts. For instance, optogenetics has enabled the dissection of the inner workings of cells, such as signaling pathway dynamics, as well as interrogation and correction of disease states in preclinical models, such as in hearing, cardiac pacing, bladder dysfunction, and cancer. In a pioneering clinical study, optogenetics restored a basic form of vision in a patient with blindness (3).

Pancreatic islet physiology relies on cell type–specific functions and the interplay between various islet (and non-islet) cell types (4). β-Cells are insulin-producing cells and essential for glucose homeostasis and metabolism (5). Glucagon-secreting α-cells exert counterregulatory effects, and somatostatin (SST)–secreting δ-cells perform paracrine inhibition of both insulin and glucagon (6,7). Islet cell type function is intimately linked to diabetes, a series of metabolic disorders characterized by chronically high blood glucose levels. In type 1 diabetes (T1D), β-cells are targeted via an autoimmune attack, whereas β-cell dysfunction and insulin resistance are characteristic for type 2 diabetes (T2D). Both forms can be managed through the administration of exogenous insulin. However, glycemic control is suboptimal and still requires challenging glucose monitoring and insulin dosing. Patients with T2D can be provided medications that remove excess glucose from the bloodstream, increase insulin sensitivity, or augment insulin secretion. For example, glucagon-like peptide 1 receptor (GLP-1R) agonists have been used to mimic the action of the incretin hormone GLP-1 (8). Yet, both forms of diabetes often still lead to a host of secondary complications, such as neuropathy, nephropathy, and retinopathy (9). It is evident that islet function, and its integration into body physiology, are both complex and dynamic, calling for experimental approaches that offer high specificity.

The past 5 years have seen a rapid increase in the number of studies that apply optogenetics to islet cells, taking advantage of its high spatiotemporal precision. These applications focused either directly on the metabolically important islet cell types in situ or “designer” replacement cells (e.g., engineered insulin-secreting cells). Here, we first introduce the fundamental working principles of optogenetics and in relation to the established field of neural optogenetics and other genetically encoded actuators deployed into islets. We then cover the form and function of light-sensitive actuators used to modulate islet cell function and signaling (Table 1), as well as effects at the cell, tissue, and organism level (Figs. 1and 2). Insulin production and secretion, cell-cell interactions, and signaling pathways are discussed, along with illumination methods (Table 2 and Fig. 3). Powerful chemogenetic and photopharmacological methods are only briefly touched upon as they have recently been reviewed elsewhere (10–13). Finally, we discuss the applicability of optogenetics in islet biology and conclude with an outlook.

Table 1

Optogenetic actuators applied to islet cells or replacement cells

ActuatorOptogeneticsystemWavelength(nm), on/offControlled processModelsReferences
CBD Glow Control 545/dark GLP-1 production Microencapsulated HEK-293 cells, db/db mice Mansouri et al. (43
CRY monSTIM1 450/dark Insulin secretion hPSCs differentiated into PIOs, STZ-treated mice Choi et al. (49
LOV-sensing domain LightOn 450/dark Insulin production Hydrodynamics-based DNA transfection, STZ-treated mice Wang et al. (27
LOV-sensing domain LuminON 450/dark Insulin production Microencapsulated HEK-293 cells, STZ-treated mice Li et al. (28
LOV-sensing domain GBOI 450/dark Insulin production Microencapsulated HEK-293 cells, STZ-treated mice Li et al. (29
LOV-sensing domain miniSOG2 450/dark Peri-islet neuronal cell ablation miniSOG2 zebrafish line Yang et al. (66
LOV-sensing domain optoPASS 450/dark Insulin secretion Microencapsulated HEK-293 cells, STZ-treated mice Wang et al. (57
PAC bPAC 450/dark Insulin secretion Adenoviral-mediated transduction of bPAC-MIN6 cells (monolayer and pseudoislets), islets of wild-type mice Zhang and Tzanakakis (53
PAC bPAC 450/dark Insulin secretion Encapsulated bPAC- MIN6 cells and pseudoislets, STZ-treated mice Zhang and Tzanakakis (54
PAC bPAC 450/dark Insulin secretion Encapsulated bPAC- EndoC-βH3 pseudoislets, STZ-treated mice Chen et al. (55
PHY BphS 730/dark Insulin or GLP-1 production Microencapsulated HEK-293 cells, STZ-treated or db/db mice Shao et al. (31), Yu et al. (32), Man et al. (33
PHY BphS 730/dark GLP-1 production Microencapsulated HEK-293 cells, db/db mice Liu et al. (44
PHY BphS 730/dark Insulin production FAID cells (engineered human mesenchymal stem cells), STZ-treated mice Yu et al. (34
PHY REDMAP 660/730 Insulin production Microencapsulated HEK-293 cells, STZ-treated mice and rats Zhou et al. (35
PHY rOpto-FGFR1 660/dark MAPK/ERK signaling INS-1E cells Reichhart et al. (79
Type 1 opsin ChR2 470/dark Insulin secretion Islets of ChR2 RIP-Cre mice, mice fed an 8-week HFD Reinbothe et al. (47
Type 1 opsin ChR2 470/dark Insulin secretion ChR2-MIN6 cells, STZ-treated mice Kushibiki et al. (48
Type 1 opsin ChR2 470/dark Islet electrical dynamics Islets of ChR2 PDX-Cre mice Westacott et al. (67
Type 1 opsin ChR2 470/dark Neural-stimulated insulin secretion ChAT-ChR2 mice Fontaine et al. (74
Type 1 opsin ChR2 470/dark Neural-stimulated insulin secretion and β-cell proliferation ChAT-ChR2 STZ-treated mice Kawana et al. (18
Type 1 opsin ChR2 470/dark α-Cell paracrine regulation Islets from ChR2 RIP-Cre mice Briant et al. (68
Type 1 opsin ChR2 470/dark δ-Cell paracrine regulation In situ SST-ChR2-YFP islets transplanted into the ACE of wild-type mice, mice fed an 8-week HFD Arrojo et al. (69
Type 1 opsin ChR2 470/dark Membrane potential and gene regulation MIN6 cells, islets of ChR2 PDX-Cre mice Miranda et al. (75
Type 1 opsin ChR2 470/dark Pericyte activity Islets from NG2-ChR2 mice transplanted into the ACE of STZ-treated mice Tamayo et al. (80
Type 1 opsin ChR2 470/dark Pericyte activity Intact but accessed pancreas of anesthetized NG2- ChR2 mice on HFD Michau et al. (81
Type 1 opsin NpHR 575/dark Silencing of β-cells Islets of NpHR INS1-Cre mice Johnston et al. (61
Type 2 opsin OPN4 470/dark GLP-1 production Microencapsulated HEK-293 cells, db/db mice Ye et al. (39
Type 2 opsin OPN4 470/dark GLP-1 production Encapsulated HeLa cells, STZ-treated mice Choi et al. (40
Type 2 opsin OPN4 470/dark GLP-1 production Microencapsulated HEK-293 cells, RIN-m5F cells, and glucose-responsive UCNPs, STZ-treated mice Lu et al. (45
Type 2 opsin OPN4 470/dark Insulin secretion Microencapsulated INSvesc cells (iβ-cells), STZ-treated mice Mansouri et al. (46
ActuatorOptogeneticsystemWavelength(nm), on/offControlled processModelsReferences
CBD Glow Control 545/dark GLP-1 production Microencapsulated HEK-293 cells, db/db mice Mansouri et al. (43
CRY monSTIM1 450/dark Insulin secretion hPSCs differentiated into PIOs, STZ-treated mice Choi et al. (49
LOV-sensing domain LightOn 450/dark Insulin production Hydrodynamics-based DNA transfection, STZ-treated mice Wang et al. (27
LOV-sensing domain LuminON 450/dark Insulin production Microencapsulated HEK-293 cells, STZ-treated mice Li et al. (28
LOV-sensing domain GBOI 450/dark Insulin production Microencapsulated HEK-293 cells, STZ-treated mice Li et al. (29
LOV-sensing domain miniSOG2 450/dark Peri-islet neuronal cell ablation miniSOG2 zebrafish line Yang et al. (66
LOV-sensing domain optoPASS 450/dark Insulin secretion Microencapsulated HEK-293 cells, STZ-treated mice Wang et al. (57
PAC bPAC 450/dark Insulin secretion Adenoviral-mediated transduction of bPAC-MIN6 cells (monolayer and pseudoislets), islets of wild-type mice Zhang and Tzanakakis (53
PAC bPAC 450/dark Insulin secretion Encapsulated bPAC- MIN6 cells and pseudoislets, STZ-treated mice Zhang and Tzanakakis (54
PAC bPAC 450/dark Insulin secretion Encapsulated bPAC- EndoC-βH3 pseudoislets, STZ-treated mice Chen et al. (55
PHY BphS 730/dark Insulin or GLP-1 production Microencapsulated HEK-293 cells, STZ-treated or db/db mice Shao et al. (31), Yu et al. (32), Man et al. (33
PHY BphS 730/dark GLP-1 production Microencapsulated HEK-293 cells, db/db mice Liu et al. (44
PHY BphS 730/dark Insulin production FAID cells (engineered human mesenchymal stem cells), STZ-treated mice Yu et al. (34
PHY REDMAP 660/730 Insulin production Microencapsulated HEK-293 cells, STZ-treated mice and rats Zhou et al. (35
PHY rOpto-FGFR1 660/dark MAPK/ERK signaling INS-1E cells Reichhart et al. (79
Type 1 opsin ChR2 470/dark Insulin secretion Islets of ChR2 RIP-Cre mice, mice fed an 8-week HFD Reinbothe et al. (47
Type 1 opsin ChR2 470/dark Insulin secretion ChR2-MIN6 cells, STZ-treated mice Kushibiki et al. (48
Type 1 opsin ChR2 470/dark Islet electrical dynamics Islets of ChR2 PDX-Cre mice Westacott et al. (67
Type 1 opsin ChR2 470/dark Neural-stimulated insulin secretion ChAT-ChR2 mice Fontaine et al. (74
Type 1 opsin ChR2 470/dark Neural-stimulated insulin secretion and β-cell proliferation ChAT-ChR2 STZ-treated mice Kawana et al. (18
Type 1 opsin ChR2 470/dark α-Cell paracrine regulation Islets from ChR2 RIP-Cre mice Briant et al. (68
Type 1 opsin ChR2 470/dark δ-Cell paracrine regulation In situ SST-ChR2-YFP islets transplanted into the ACE of wild-type mice, mice fed an 8-week HFD Arrojo et al. (69
Type 1 opsin ChR2 470/dark Membrane potential and gene regulation MIN6 cells, islets of ChR2 PDX-Cre mice Miranda et al. (75
Type 1 opsin ChR2 470/dark Pericyte activity Islets from NG2-ChR2 mice transplanted into the ACE of STZ-treated mice Tamayo et al. (80
Type 1 opsin ChR2 470/dark Pericyte activity Intact but accessed pancreas of anesthetized NG2- ChR2 mice on HFD Michau et al. (81
Type 1 opsin NpHR 575/dark Silencing of β-cells Islets of NpHR INS1-Cre mice Johnston et al. (61
Type 2 opsin OPN4 470/dark GLP-1 production Microencapsulated HEK-293 cells, db/db mice Ye et al. (39
Type 2 opsin OPN4 470/dark GLP-1 production Encapsulated HeLa cells, STZ-treated mice Choi et al. (40
Type 2 opsin OPN4 470/dark GLP-1 production Microencapsulated HEK-293 cells, RIN-m5F cells, and glucose-responsive UCNPs, STZ-treated mice Lu et al. (45
Type 2 opsin OPN4 470/dark Insulin secretion Microencapsulated INSvesc cells (iβ-cells), STZ-treated mice Mansouri et al. (46

ChAT, choline acetyltransferase promoter; MAPK/ERK, mitogen-activated protein kinases/extracellular signal–regulated kinase; NG2, nerve/glial antigen 2; PAC, photoactivatable adenylyl cyclase; PDX, pancreatic and duodenal homeobox; PHY, phytochrome; RIP, rat insulin promoter.

Figure 1

Optogenetic actuators to control the transcription of GLP-1 and insulin. A: LightOn employs the synthetic transcription factor GAVPO that contains Gal4, VVD, and p65. VVD dimerizes in the presence of blue light, allowing for Gal4 to bind galactose-responsive upstream activation sequence (UASG) elements to drive transcription. B: GBOI is a two-component system whereby the GI-Gal4 fusion protein is under the control of the glucose-sensitive promoter PGIP. The constitutively expressed LOV-VP16 forms a complex with GI-Gal4 only in the presence of high glucose and blue light to drive insulin expression. C: The FRL-activated BphS system converts GTP to c-di-GMP, which binds to FRTA (which contains the BldD transcription factor, VP64, and p65). Activated FRTA subsequently drives transcription from BldD-specific DNA operator sites (PFRL). Coexpression of the c-di-GMP phosphodiesterase YhjH can be used to reduce basal intracellular c-di-GMP. D: REDMAP uses a fused ΔphyA-Gal4 construct along with FHY1 fused to a VP64 transactivator. The ΔphyA-FHY1 interaction is promoted by red light to mediate transgene expression. The system can be reversed with FRL. E: OPN4 can drive endogenous PLC signaling, Ca2+ influx, and NFAT cell–mediated transcription. F: The Glow Control system uses a membrane-anchored TtCBD and a second TtCBD domain fused to VP64, p65, and Rta (VPR) and TetR. In the dark, TtCBD domains are fused. Green light promotes dissociation, leading to transgene expression. hGLP1, human glucagon-like peptide 1; TetO, tetracycline operator.

Figure 1

Optogenetic actuators to control the transcription of GLP-1 and insulin. A: LightOn employs the synthetic transcription factor GAVPO that contains Gal4, VVD, and p65. VVD dimerizes in the presence of blue light, allowing for Gal4 to bind galactose-responsive upstream activation sequence (UASG) elements to drive transcription. B: GBOI is a two-component system whereby the GI-Gal4 fusion protein is under the control of the glucose-sensitive promoter PGIP. The constitutively expressed LOV-VP16 forms a complex with GI-Gal4 only in the presence of high glucose and blue light to drive insulin expression. C: The FRL-activated BphS system converts GTP to c-di-GMP, which binds to FRTA (which contains the BldD transcription factor, VP64, and p65). Activated FRTA subsequently drives transcription from BldD-specific DNA operator sites (PFRL). Coexpression of the c-di-GMP phosphodiesterase YhjH can be used to reduce basal intracellular c-di-GMP. D: REDMAP uses a fused ΔphyA-Gal4 construct along with FHY1 fused to a VP64 transactivator. The ΔphyA-FHY1 interaction is promoted by red light to mediate transgene expression. The system can be reversed with FRL. E: OPN4 can drive endogenous PLC signaling, Ca2+ influx, and NFAT cell–mediated transcription. F: The Glow Control system uses a membrane-anchored TtCBD and a second TtCBD domain fused to VP64, p65, and Rta (VPR) and TetR. In the dark, TtCBD domains are fused. Green light promotes dissociation, leading to transgene expression. hGLP1, human glucagon-like peptide 1; TetO, tetracycline operator.

Close modal
Figure 2

Optogenetic actuators deployed into islet cells to control intracellular signaling, Ca2+ influx, and insulin secretion. A: Native Gq-coupled GPCRs promote PLC signaling to raise cytoplasmic Ca2+ concentrations and potentiate insulin secretion. OPN4 can be used to achieve light-dependent control of this process. B: The light-sensitive ion channel ChR2 can be used to promote membrane depolarization (not shown) and influx of cations such as Ca2+. C: Ca2+-specific influx can be achieved through activation of native CRACs using light-sensitive monSTIM1. High intracellular Ca2+ stimulates insulin production and exocytosis of insulin from secretory granules. D: AC activity can enhance insulin secretion through the cAMP/PKA pathway, and this can be optically controlled with bPAC. The native GLP-1 pathway is also shown here. E: Secretion can also be light-regulated at the post-transcriptional level from the ER using optoPASS. F: Finally, the anion pump, NpHR inhibits insulin secretion in a light-dependent manner by promoting the influx of Cl into the cell. Dotted lines indicate indirect insulin release. DAG, diacylglycerol; EPAC, exchange protein activated by cAMP; IP3, inositol triphosphate; PIP2, phosphatidylinositol 4,5-bisphosphate; VGCC, voltage-gated Ca2+ channel.

Figure 2

Optogenetic actuators deployed into islet cells to control intracellular signaling, Ca2+ influx, and insulin secretion. A: Native Gq-coupled GPCRs promote PLC signaling to raise cytoplasmic Ca2+ concentrations and potentiate insulin secretion. OPN4 can be used to achieve light-dependent control of this process. B: The light-sensitive ion channel ChR2 can be used to promote membrane depolarization (not shown) and influx of cations such as Ca2+. C: Ca2+-specific influx can be achieved through activation of native CRACs using light-sensitive monSTIM1. High intracellular Ca2+ stimulates insulin production and exocytosis of insulin from secretory granules. D: AC activity can enhance insulin secretion through the cAMP/PKA pathway, and this can be optically controlled with bPAC. The native GLP-1 pathway is also shown here. E: Secretion can also be light-regulated at the post-transcriptional level from the ER using optoPASS. F: Finally, the anion pump, NpHR inhibits insulin secretion in a light-dependent manner by promoting the influx of Cl into the cell. Dotted lines indicate indirect insulin release. DAG, diacylglycerol; EPAC, exchange protein activated by cAMP; IP3, inositol triphosphate; PIP2, phosphatidylinositol 4,5-bisphosphate; VGCC, voltage-gated Ca2+ channel.

Close modal
Table 2

Light sources applied in optogenetic islet and diabetes studies (cell lines, primary islets, and rodent models)

Device/light sourceApplicationDeliverymethodIllumination areaKey features and benefitsCommonactuators usedReferences
Microscope(e.g., wide field, confocal, or two photon) In vitro, ex vivo LED or laser Subcellular, single cell, cell population High spatial resolution, can couple with imaging modalities and biosensors, minimal or no specialized hardware required ChR2, NpHR Johnston et al. (61), Westacott et al. (67), Briant et al. (68), Arrojo et al. (69), Miranda et al. (75
Microscope(e.g., wide field, confocal, or two photon) In situ implantation into ACE LED or laser Subcellular, single cell, cell population High spatial resolution, can couple with imaging modalities and biosensors, minimal or no specialized hardware required ChR2, NpHR Arrojo et al. (69
LED arrays, matrixes, or strips In vitro, ex vivo Placed above or below tissue culture plates Cell ensemble Single-well control (in some instances), multicolor stimulation, low cost CBD, ChR2, CRY, OPN4, PAC, PHY Shao et al. (31), Yu et al. (34), Zhou et al. (35), Ye et al. (39), Mansouri et al. (43), Mansouri et al. (46), Choi et al. (49), Zhang and Tzanakakis (53), Chen et al. (55
LED arrays, matrixes, or strips In vivo designer cell implantation Whole-cage illumination(e.g., above or surrounding the cage) Organism Noninvasive, tether-free, low cost CBD, ChR2, CRY, LOV, OPN4, PAC, PHY Li et al. (29), Zhou et al. (35), Ye et al. (39), Mansouri et al. (43), Mansouri et al. (46), Choi et al. (49), Zhang and Tzanakakis (54), Chen et al. (55
LED arrays, matrixes, or strips In situ islet implantation into ACE Whole-cage illumination(e.g., above or surrounding the cage) Organism Noninvasive, tether-free, low cost CBD, ChR2, CRY, LOV, OPN4, PAC, PHY Tamayo et al. (80
External laser In vivo UCNP stimulation (pancreas or designer cell implant) External animal illumination (e.g., above the cage) Organism Noninvasive, tether-free, suitable for long-term illumination ChR2, OPN4 Kawana et al. (18), Lu et al. (45
Lamp In vitro, ex vivo Culture plate illumination Cell population Noninvasive, tether-free, suitable for long-term illumination LOV Wang et al. (27), Li et al. (28
Lamp Designer cell implantation in vivo Whole-cage illumination Organism Noninvasive, tether-free, suitable for long-term illumination PHY Shao et al. (31), Man et al. (33), Yu et al. (34
Optical fiber or cannula In vitro, ex vivo Placed above, below, or inside tissue culture plates or microtubes Cell ensemble Directed to specific area or tissue, can be coupled with hydrogels ChR2 Reinbothe et al. (47), Kushibiki et al. (48), Briant et al. (68
Optical fiber or cannula In vivo light delivery to specific area or tissue Implant Tissue or cell population Directed to specific area or tissue, can be coupled with hydrogels ChR2, LOV Kawana et al. (18), Wang et al. (27), Fontaine et al. (74), Michau et al. (81
Optical fiber or cannula Coencapsulation with designer cells for implantation in vivo Implant Tissue or cell population Directed to specific area or tissue, can be coupled with hydrogels ChR2, OPN4 Ye et al. (39), Choi et al. (40), Kushibiki et al. (48
Wireless LED implant or patch Designer cell implantation in vivo Implant Tissue or cell population Tether-free, suitable for long-term illumination, can be coupled with hydrogels CBD, PHY Shao et al. (31), Yu et al. (32), Mansouri et al. (43), Liu et al. (44
Device/light sourceApplicationDeliverymethodIllumination areaKey features and benefitsCommonactuators usedReferences
Microscope(e.g., wide field, confocal, or two photon) In vitro, ex vivo LED or laser Subcellular, single cell, cell population High spatial resolution, can couple with imaging modalities and biosensors, minimal or no specialized hardware required ChR2, NpHR Johnston et al. (61), Westacott et al. (67), Briant et al. (68), Arrojo et al. (69), Miranda et al. (75
Microscope(e.g., wide field, confocal, or two photon) In situ implantation into ACE LED or laser Subcellular, single cell, cell population High spatial resolution, can couple with imaging modalities and biosensors, minimal or no specialized hardware required ChR2, NpHR Arrojo et al. (69
LED arrays, matrixes, or strips In vitro, ex vivo Placed above or below tissue culture plates Cell ensemble Single-well control (in some instances), multicolor stimulation, low cost CBD, ChR2, CRY, OPN4, PAC, PHY Shao et al. (31), Yu et al. (34), Zhou et al. (35), Ye et al. (39), Mansouri et al. (43), Mansouri et al. (46), Choi et al. (49), Zhang and Tzanakakis (53), Chen et al. (55
LED arrays, matrixes, or strips In vivo designer cell implantation Whole-cage illumination(e.g., above or surrounding the cage) Organism Noninvasive, tether-free, low cost CBD, ChR2, CRY, LOV, OPN4, PAC, PHY Li et al. (29), Zhou et al. (35), Ye et al. (39), Mansouri et al. (43), Mansouri et al. (46), Choi et al. (49), Zhang and Tzanakakis (54), Chen et al. (55
LED arrays, matrixes, or strips In situ islet implantation into ACE Whole-cage illumination(e.g., above or surrounding the cage) Organism Noninvasive, tether-free, low cost CBD, ChR2, CRY, LOV, OPN4, PAC, PHY Tamayo et al. (80
External laser In vivo UCNP stimulation (pancreas or designer cell implant) External animal illumination (e.g., above the cage) Organism Noninvasive, tether-free, suitable for long-term illumination ChR2, OPN4 Kawana et al. (18), Lu et al. (45
Lamp In vitro, ex vivo Culture plate illumination Cell population Noninvasive, tether-free, suitable for long-term illumination LOV Wang et al. (27), Li et al. (28
Lamp Designer cell implantation in vivo Whole-cage illumination Organism Noninvasive, tether-free, suitable for long-term illumination PHY Shao et al. (31), Man et al. (33), Yu et al. (34
Optical fiber or cannula In vitro, ex vivo Placed above, below, or inside tissue culture plates or microtubes Cell ensemble Directed to specific area or tissue, can be coupled with hydrogels ChR2 Reinbothe et al. (47), Kushibiki et al. (48), Briant et al. (68
Optical fiber or cannula In vivo light delivery to specific area or tissue Implant Tissue or cell population Directed to specific area or tissue, can be coupled with hydrogels ChR2, LOV Kawana et al. (18), Wang et al. (27), Fontaine et al. (74), Michau et al. (81
Optical fiber or cannula Coencapsulation with designer cells for implantation in vivo Implant Tissue or cell population Directed to specific area or tissue, can be coupled with hydrogels ChR2, OPN4 Ye et al. (39), Choi et al. (40), Kushibiki et al. (48
Wireless LED implant or patch Designer cell implantation in vivo Implant Tissue or cell population Tether-free, suitable for long-term illumination, can be coupled with hydrogels CBD, PHY Shao et al. (31), Yu et al. (32), Mansouri et al. (43), Liu et al. (44

PAC, photoactivatable adenylyl cyclase; PHY, phytochrome.

Figure 3

Illumination strategies applied in vivo (rodent models). A: Animals can be stimulated using LED lamps, matrixes, or arrays installed above, around, or within the animal cage; these illuminate multiple animals (left). Direct local stimulation to the pancreas or a cell implant can be achieved using tethered optical fibers (right). B: Wireless LED implants allow for tether-free light delivery in freely behaving animals. They can be coencapsulated with designer cells, as represented here, or be used to directly stimulate a tissue of interest. Wireless LEDs are powered through a field generator wire surrounding the cage. C: UCNPs achieve minimally invasive tether-free light delivery deep into tissues by converting NIR light into higher-energy visible light. This is particularly useful for actuators that require short wavelengths or for tissues that are difficult to target with external illumination strategies. Animals can be stimulated externally with NIR - typically using LED lamps, arrays, or matrixes as illustrated in A. UCNPs can be injected into the pancreas for direct tissue stimulation or coencapsulated with designer cells for transplantation.

Figure 3

Illumination strategies applied in vivo (rodent models). A: Animals can be stimulated using LED lamps, matrixes, or arrays installed above, around, or within the animal cage; these illuminate multiple animals (left). Direct local stimulation to the pancreas or a cell implant can be achieved using tethered optical fibers (right). B: Wireless LED implants allow for tether-free light delivery in freely behaving animals. They can be coencapsulated with designer cells, as represented here, or be used to directly stimulate a tissue of interest. Wireless LEDs are powered through a field generator wire surrounding the cage. C: UCNPs achieve minimally invasive tether-free light delivery deep into tissues by converting NIR light into higher-energy visible light. This is particularly useful for actuators that require short wavelengths or for tissues that are difficult to target with external illumination strategies. Animals can be stimulated externally with NIR - typically using LED lamps, arrays, or matrixes as illustrated in A. UCNPs can be injected into the pancreas for direct tissue stimulation or coencapsulated with designer cells for transplantation.

Close modal

As optogenetic methods rely on a physical stimulation modality (light), they have the potential to overcome some of the limitations inherent to pharmacological or genetic methods. First, unlike small molecules or biologics, light can be applied remotely (i.e., through tissue barriers to some depths [see below]) and withdrawn easily (i.e., without retention in the cellular milieu). Second, unlike genetic modification, signals can be precisely manipulated on very short (e.g., even down to milliseconds) or much longer timescales (e.g., up to days or weeks). Third, and rather uniquely, spatial precision can be achieved by delivering light to only the desired tissues, cells, or even subcellular compartments, reducing off-target effects. Fourth, in many optogenetic experiments, biological systems are studied in their intact state, with spatial information and signaling dynamics being preserved.

Optogenetics is now a staple technique in neuroscience primarily for dissecting brain circuits and behavior (14). There are parallels between applications to islets and neurons. For instance, on the tissue level, it is difficult, but desirable, to decipher the role of individual cell populations (e.g., one specific type of neuron or neurons vs. glia, compared with β-cells vs. α-cells in islets). This commonality has inspired the transfer of transgenic models between the fields. On the cellular level, the presence of voltage-gated ion channels endow β-, δ-, and α-cells with electrical excitability (15). This provides opportunity to control Ca2+ influx and hormone secretion dynamics using the same light-sensitive proteins commonly used to stimulate (16) or inhibit (17) neurons. Notably, optogenetic neural manipulation has been used to study islet-specific effects of neural-endocrine cell interactions (18). Thus, neuronal optogenetics has both informed and become integrated into islet biology.

Optogenetics was not the first neuroscience-derived method applied to islet cells. Chemogenetic systems, most prominently designer receptors exclusively activated by designer drugs (DREADDs), achieve orthogonal and temporal control over G-protein–coupled receptor (GPCR) signaling pathways and thereby changes in cell and organism state (19,20). In lieu of their endogenous ligand, DREADDs are engineered to bind a pharmacologically inert drug. DREADDs have been applied in islets to study, for instance, GPCR-mediated signaling and endocrine cell function (13,21). Both DREADDs and optogenetic actuators permit interrogation of cellular processes with some spatial precision through selective expression in transgenic mouse lines or localized viral delivery. However, optogenetics can provide further specificity through confined illumination. The timescales accessible also differ, with chemogenetics suited for medium- or long-term chronic stimulation (i.e., hours to weeks) and optogenetics for shorter stimulation (i.e., milliseconds to minutes, or longer as required) and greater flexibility in stimulation protocols (e.g., pulsed vs. continuous illumination). In terms of stimulus delivery, chemogenetics negates the need for optical devices, and light penetration deep into tissues can be limiting (see below). However, in vivo models, such as islet transplantation into the anterior chamber of the eye (ACE) (22,23), provide light accessibility, and the use of upconversion nanoparticles (UCNPs) achieve minimally invasive, tether-free light delivery deep into tissues (24,25). Finally, most current ligand-orthogonal actuators are GPCRs and ion channels, whereas the optogenetic toolkit has great diversity to control a variety of processes. Thus, the researcher’s choice will largely depend on the process being investigated and the properties of available tools.

Insulin Production

Insulin is natively expressed from the INS locus on human chromosome 11 under the control of the highly β-cell–specific insulin promoter. It is first synthesized as a single-chain (pre) proinsulin before being processed into granules and secreted via exocytosis (26). Several light-dependent transcription systems have been developed to produce insulin using transplanted designer cells in animal models. Most of these studies delivered light sources in the cage or through fiber optic implants (Fig. 3A). The LightOn system employs a light-sensitive chimeric transcription factor termed GAVPO that contains the light-oxygen-voltage (LOV)–sensing domain of the fungal protein vivid (VVD), the DNA-binding domain of the transcription factor Gal4, and a p65 transactivation domain (27) (Fig. 1A). Blue light promotes VVD dimerization, allowing Gal4 to bind to galactose-responsive upstream activation sequence (UAS) elements that drive insulin expression on a second locus. Wang et al. (27) delivered this two-component system into streptozotocin (STZ)-treated T1D mice via hydrodynamic plasmid transfection and observed reduced blood glucose levels after 8 h of illumination. More recently, the luminescent version LuminON was generated (28). This facultatively bioluminescence resonance energy transfer–based system contains nanoluciferase (nLuc) fused to GAVPO and can be controlled either by light or the luciferase substrate furimazine. Pulsatile expression of the hormone was achieved in microencapsulated engineered LuminON human embryonic kidney 293 (HEK-293) cells transplanted into STZ-treated T1D mice. However, illumination resulted in shortened antidiabetic efficacy compared with furimazine, likely attributed to poor penetration of blue light.

GBOI is the glucose-blue light chemi-optogenetic cell–implanted therapy system in which insulin production is only triggered in the presence of light and high blood glucose (29) (Fig. 1B). The system harnesses a LOV domain from Arabidopsis thaliana (30) and consists of a LOV-VP16 fusion protein constitutively expressed from the cytomegalovirus (CMV) promoter PCMV-LOV-VP16, a Gigantea (GI)-Gal4 fusion construct under the control of the glucose-dependent insulinotropic polypeptide (GIP) promoter PGIP-GI-Gal4, and human preproinsulin under control of the Gal4-responsive UAS promoter PUAS-insulin. High glucose induces GI-Gal4 expression, and in the presence of blue light, LOV-VP16 and GI-Gal4 form a complex to drive insulin expression. STZ-treated T1D mice subcutaneously implanted with microencapsulated GBOI HEK-293 cells displayed a reduction in blood glucose and increased serum insulin. Glucose tolerance tests (GTTs) under continuous illumination revealed that GBOI-implanted animals also had lower postprandial glucose levels. The GBOI system could be programmed into already-existing blood glucose monitoring devices that would only illuminate cells when blood glucose levels exceed a threshold.

The limited tissue penetration of blue light (e.g., less than a few millimeters for 450–480 nm light of a moderate nontoxic intensity) has prompted the development of actuators sensitive to longer light wavelengths, in particular red light. Reduced light-induced cytotoxicity during long-term illumination is an added advantage of red light stimulation. A far-red light (FRL)–controlled system was engineered to drive insulin expression using the bacterial diguanylate cyclase, BphS (31) (Fig. 1C). Activation with FRL promotes BphS to convert GTP to cyclic diguanylate monophosphate (c-di-GMP), which then binds to a synthetic transactivating element (FRTA) containing the Streptomyces coelicolor–derived transcription factor, BldD. C-di-GMP–promoted dimerization of FRTA leads to its binding to BldD-specific DNA operator sites upstream of insulin. HEK-293 cells containing these elements were subcutaneously implanted in STZ-treated T1D mice. FRL illumination elicited improved blood glucose profiles and tolerance. It also resulted in sustained and detectable levels of insulin in the bloodstream. These genetic components were further combined with a semiautomatic illumination system, comprising smartphone technology, glucose sensing, and a wirelessly powered FRL micro–light-emitting diode (LED) integrated with the cells in a hydrogel (31,32) (Fig. 3B). Most recently, an updated semiautonomous closed-loop bioelectronic system was developed (33) that achieved light-mediated insulin expression for up to 1 month in STZ-treated T1D mice.

The BphS system has also been tested in FRL-activated human islet–like designer (FAID) cells (34). BphS was stably integrated into human mesenchymal stem cells, which are commonly explored in cell therapy. Encapsulated subcutaneous transplantations in STZ-treated T1D mice were illuminated for up to 40 days. This is likely the longest experimental timeline to date using a light-induced engineered insulin-producing cell therapy. FRL illumination improved glucose tolerance and homeostasis, increased weight gain compared with diabetic controls, and improved polyuria. Most importantly, monitoring of nonfasting blood glucose demonstrated that illuminated STZ-FAID mice had fewer fluctuations compared with diabetic mice administered insulin. Lowered hemoglobin A1c levels were also achieved, indicating longer-term blood glucose control. Interestingly, assessment of diabetes-related complications revealed amelioration of fibrosis as well as renal and cardiac damage.

Another recent red light/FRL-sensitive approach is the miniaturized Δphytochrome A (ΔphyA)–based photoswitch (REDMAP) (35) (Fig. 1D). PhyA is a plant photoreceptor that also utilizes phycocyanobilin as its chromophore. Red light/FRL promotes/reverses nuclear translocation of phyA through interactions with FHY1 and FHY1-like shuttle proteins (36,37). Zhou et al. (35) generated a transcriptional activation system by fusing the truncated photoreceptor to Gal4 (ΔphyA-Gal4) and FHY1 to a VP64 transactivator (FHY1-VP64). Illumination with red light promoted the ΔphyA-FHY1 interaction, nuclear translocation, and transcriptional activation of a target gene downstream of synthetic UAS repeats. For gene inactivation, FRL illumination disrupts the ΔphyA-FHY1 interaction, leading to dissociation of ΔphyA-Gal4 and FHY1-VP64. The system was applied in vivo by transplanting encapsulated REDMAP HEK-293 cells into STZ-treated T1D mice to achieve light-dependent insulin production. Insulin levels were increased along with a reduction in blood glucose levels. Moreover, GTTs revealed that a single pulse of red light was sufficient to elicit substantial improvement in glucose homeostasis. Studies over weeks also showed improved glucose tolerance and reduced blood glucose levels. In contrast to BphS, REDMAP displays efficient and high transcriptional activation (>150-fold) along with rapid (de)activation kinetics (∼1 s compared with hours).

GLP-1 Production

Insulin secretion can be further enhanced through incretin hormones, such as GLP-1, which targets the highly expressed GLP-1R (38). Several groups have generated designer cells that produce and release GLP-1 in a light-dependent manner. The earliest approach was by coexpression of melanopsin (OPN4), a blue light–activated GPCR of the opsin family, and a short variant of GLP-1 (shGLP-1) under control of the nuclear factor–activated T cell (NFAT) promoter (39) (Fig. 1E). OPN4 activation relies on photoisomerization of the chromophore 11-cis retinal, a form of vitamin A (39). When engineered into HEK-293 cells, activated OPN4 drives endogenous phospholipase C (PLC) signaling, leading to Ca2+ influx and NFAT-mediated transcription of shGLP-1. This synthetic transcription device combined heterologous and endogenous factors in designer HEK-293 cells and was applied in db/db T2D mice (39). Pulsed blue light illumination increased blood insulin levels and reduced glycemic excursions. The system has also been combined with a light-guiding hydrogel implant (40).

A complementary system termed Glow Control used green light–sensitive cobalamin-binding domains (CBDs) from the Thermus thermophilus CarH protein (TtCBD) (Fig. 1F). These domains bind the 5′-deoxyadenosylcobalamin chromophore, an active vitamin B12 coenzyme, and form tetrameric assemblies in the dark that dissociate into monomers (41,42). Mansouri et al. (43) engineered GlowhGLP1 HEK-293 cells that expressed membrane-anchored TtCBD as well as a transactivator containing a synthetic transcription factor (VP64, p65, and the transcriptional activator, Rta) fused with a tetracycline repressor (TetR) and TtCBD. Expression of a target gene (e.g., GLP-1) is achieved using synthetic tetracycline-responsive promoter elements. In the dark, the transactivator is sequestered at the plasma membrane via TtCBD oligomerization. Green light promotes transactivator release and TetR binding to the tetracycline-responsive promoter to drive gene expression. Light-dependent GLP-1 production was achieved using subcutaneous implantation in a db/db T2D mouse model. Illumination was performed using an implant that mimicked the green light emitted from smart watches and resulted in lowered fasting glucose levels, attenuated insulin resistance, and reduced body weight by day 12 of the illumination period. However, protection of such sensitive implants from ambient light was highlighted as an important consideration for efficacy.

In addition to insulin production, the BphS FRL system was also used to drive shGLP-1 transcription in Shao et al. using db/db T2D mice, resulting in insulin tolerance, reduced insulin resistance, and restored glucose homeostasis (31). As a further development, a self-powered optogenetic system was established (44). The device converts biomechanical energy from respiration movement in the animal into electricity to power the LED implant. A db/db T2D mouse model illuminated with the self-powered optogenetic system showed fast regeneration of blood glucose levels, insulin tolerance, and blood glucose homeostasis.

Another recent advance has involved coupling glucose levels to illumination control using specialized UCNPs. These are a class of nanomaterials that can convert near infrared (NIR) light into higher-energy visible light. Specialized glucose-sensitive UCNPs have been generated to adaptively tune the intensity of nanoparticle luminescence to blood glucose concentrations (45). Lu et al. (45) demonstrated their applicability in closed-loop blood glucose homeostasis. Glucose-sensing UCNPs were encapsulated in a hydrogel with OPN4 GLP-1 HEK-293 cells (Fig. 3C). The RIN-m5F murine islet line was also included in the hydrogel to accomplish GLP-1–stimulated insulin release in an STZ-treated T1D mouse model.

Insulin Secretion

The insulin secretory pathway is regulated by glucose and modulatory ligands, such as fatty acids, neurotransmitters, and incretin hormones. As β-cells metabolize glucose the elevated ATP/ADP ratio causes the block of KATP channels. The resulting depolarization triggers the influx of extracellular Ca2+ across the plasma membrane through voltage-gated Ca2+ channels, ultimately leading to exocytosis of insulin via secretory granules (6). Additional pathways can enhance insulin release, such as the PLC and cAMP/protein kinase A (PKA) pathways. A range of optogenetic approaches have been used to control these second messengers and have mostly combined designer cell implantation with light sources in the cage or fiber optic implants (Fig. 3A). An OPN4-based approach has been employed to promote rapid Ca2+-mediated insulin release from β-cell secretory granules (46) (Fig. 2A). The human INSvesc cell line was engineered to express OPN4 to create designer iβ-cells. INSvesc cells are recalcitrant to glucose levels and contain a constitutively expressed proinsulin-nLuc. Activation of OPN4 and downstream intracellular Ca2+ signals lead to the rapid on-demand release of proinsulin-nLuc from preloaded vesicular granules. In vitro, this system achieved insulin secretion that more closely mimics native minute timescales compared with the OPN4 transcription-based system (39). Subcutaneous transplantation of microencapsulated iβ-cells into a STZ-treated T1D mouse model revealed that smartphone white light was sufficient to increase blood insulin levels and restore glucose homeostasis.

The influx of Ca2+ into β-cells has also been manipulated more directly. The blue light–sensitive ion channel channelrhodopsin-2 (ChR2) derived from Chlamydomonas algae is widely applied in neuronal optogenetics. Opening of ChR2 promotes influx of cations, including Ca2+, and consequently, cell depolarization (16) (Fig. 2B). Reinbothe et al. (47) generated a transgenic mouse model expressing ChR2 under control of the insulin promoter to stimulate insulin secretion in β-cells. Patch-clamp analysis of dispersed islets revealed a rapid electrical response to blue light in ChR2-expressing β-cells. Glucose titrations demonstrated light-induced insulin secretion at a range of concentrations in intact islets; however, further potentiation was not observed at high glucose levels, likely due to an already highly stimulated secretory pathway. Interestingly, islets from mice exposed to a high-fat diet (HFD) displayed higher levels of light-dependent insulin secretion compared with control animals, suggesting the involvement of potentiation mechanisms. In a related study, Kushibiki et al. (48) generated ChR2-expressing MIN6 β-cells for use in transplantation experiments. Insulin secretion could be triggered in the absence of glucose in vitro, and subcutaneous engraftment of these cells into STZ-treated T1D mice resulted in a reduction of blood glucose levels.

Although ChR2 activation has been used to induce secretion, its pore is not Ca2+ selective (16). Most recently, Choi et al. (49) demonstrated light-dependent regulation of the Ca2+ release–activated channel (CRAC) Orai1. The authors used monster-opto-stromal interaction molecule 1 (monSTIM1), an optimized, highly blue light–sensitive variant of OptoSTIM1 (50), to regulate intracellular Ca2+. This actuator comprises a truncated STIM1 fused to GFP and the photolyase homology region domain of the flavin-binding plant photoreceptor cryptochrome 2 (CRY2) (51). Light-induced oligomerization of monSTIM1 (mediated through CRY2) leads to activation of endogenous CRAC and Ca2+ influx (52) (Fig. 2C). The monSTIM1 construct was introduced into human pluripotent stem cells (hPSCs) via a CRISPR-Cas9–mediated genomic knock-in approach. Rapid, reversible, and repetitive Ca2+ influx was achieved both before and after cells were differentiated into pancreatic islet–like organoids (PIOs). Insulin secretion was also demonstrated in STZ-treated T1D mice subcutaneously transplanted with monSTIM1 PIOs. Illumination promoted reduction in GTTs and C-peptide levels 3–4 days after transplantation. However, illuminated monSTIM1 PIOs presented delayed insulin secretion under high-glucose conditions. The underlying mechanism is unknown, but endogenous STIM1 and monSTIM1 may be working competitively in CRAC activation.

Adenylyl cyclase (AC) downstream of GLP-1R synthesizes cAMP from ATP and thereby activates PKA, leading to enhanced Ca2+ influx and secretion of insulin (Fig. 2D). A blue light–sensitive bacterial photoactivatable AC (bPAC) from Beggiatoa has been applied to regulate cAMP signaling in β-cells. In Zhang and Tzanakakis (53), illumination of an engineered MIN6 β-cell line showed cAMP elevation and enhanced insulin release on the scale of minutes (thus on similar timescales as β-cells when exposed to varying glucose levels). Notably, the actuator was comparable to known secretagogues, such as forskolin, and its action was Ca2+-dependent. Similar effects were observed in virally transduced pseudoislets formed from MIN6 monolayer cultures and murine islets. Insulin secretion was augmented under both low- and high-glucose conditions; however, in the absence of glucose, illuminated bPAC-MIN6 cells exhibited increased cAMP but unchanged insulin secretion. Thus, this approach is different from ChR2 activation, which can stimulate insulin secretion in the absence of glucose and may pose a potential risk of hypoglycemic events. In a follow-up study, Zhang and Tzanakakis (54) transplanted MIN6-bPAC–engineered pseudoislets in an STZ-treated T1D mouse model. Animals displayed improved glucose tolerance, lower hyperglycemia, and greater plasma insulin concentrations. Importantly, the number of engineered bPAC-expressing cells required for transplantation was 2.5 times lower than in the case of control pseudoislets, demonstrating the potential for cell therapy with reduced cell numbers. More recently, the bPAC system was applied to engineered human EndoC-βH3 pseudoislets transplanted into STZ-treated T1D mice (55). However, in both in vivo studies, euglycemia was not achieved, which was attributed to the reduced viability of the cells 3–5 days post-transplantation.

Most recently, synthetic protease–based platforms enabled rapid secretion of proteins in response to light (56,57). For example, the protein of interest is fused to an endoplasmic reticulum (ER) retention motif, synthetic transmembrane domain, furin cleavage site, and protease cleavage site. This strategy allows for localization of pretranslated synthetic insulin within the ER membrane. The second key component is a light-sensitive split protease in which each half of the enzyme is fused to a specific light-sensitive magnet domain (positive magnet or negative magnet) (58). These engineered LOV domains heterodimerize in the presence of blue light to reconstitute the functional protease leading to protein release from the ER and rapid on-demand secretion (Fig. 2E). Wang et al. (57) applied their optical protease-based rapid protein secretion system (optoPASS) to secrete insulin in a STZ-treated T1D mouse model. Subcutaneous transplantation of engineered microencapsulated optoPASSinsulin HEK-293 cells and acute blue light illumination led to an increase in serum insulin levels by 15 min and resulted in nondiabetic glucose levels after 1 h.

Interactions Between Islet Cells

Islets are multicellular microorgans that exhibit paracrine signaling. As mentioned above, δ-cells produce and release SST, which inhibits glucagon secretion from α-cells and insulin from β-cells. Insulin is also an inhibitor of glucagon secretion. Along with paracrine interactions, β-cell-to-β-cell interactions occur through cell adhesion molecules (59). Optogenetic actuators have been applied to dissect these complex relationships in intact islets with preserved spatial organization and temporal interaction dynamics. These studies typically used transgenic mouse models combined with microscopy or fiber-optic stimulation (Table 2 and Fig. 3A).

Individual β-cells within an islet display significant variability in several properties, including metabolic activity, electrical dynamics, insulin secretion, and glucose sensitivity (60). However, little is known about how this cell heterogeneity impacts islet function. β-Cells are electrically coupled via connexin 36 gap junctions, which enable highly coordinated and synchronized Ca2+ waves across intact islets in response to glucose. Johnston et al. (61) used high-speed Ca2+ imaging in whole islets to find that rare and highly connected β-cell subregions exist (termed hubs). In their firing, hubs repeatedly both advance and outlast the general β-cell population. In an elegant follow-up experiment (61), hubs were interrogated by optogenetic hyperpolarization and silencing of individual cells in islets from transgenic mice expressing an enhanced version of the yellow light–sensitive chloride (Cl) pump halorhodopsin (NpHR) (17) (Fig. 2F). Strikingly, intraislet responses to high-glucose conditions as well as insulin release were disrupted upon silencing of hubs. Further characterization revealed that hub cells are highly metabolic and contain features of both mature and immature β-cells. Finally, islets were challenged with cytokines, which are implicated in both forms of diabetes and often used in cell models to mimic proinflammatory conditions and trigger islet dysfunction. This resulted in a reduced number of hubs and their connections, suggesting that hub failure may contribute to disease pathogenesis. In addition to NpHRs, several other light-gated ion channels and pumps have been used to inhibit neuronal activity and could potentially be applied to endocrine cells (62–64). Alternatively, silencing or inhibition can be achieved with methods that produce reactive oxygen species upon illumination, leading to cell ablation (65). This was recently demonstrated with the LOV domain–containing small fluorescent mini singlet oxygen generator 2 (miniSOG2) in the peri-islets of zebrafish (66).

In lieu of silencing, Westacott et al. (67) stimulated different proportions of islets in a murine β-cell–specific-ChR2 expression model to understand how subpopulations control Ca2+ responses and insulin secretion. Visualization of these responses identified islet subregions that more readily undergo light-dependent depolarization and propagate Ca2+ signals. Furthermore, specific regions assume a pacemaker-like role in initiating Ca2+ wave propagation. Taken together, these optogenetic inhibition and activation studies demonstrate how precise manipulation of Ca2+ dynamics in spatially defined cells can reveal islet-level functional interactions emerging from β-cell subpopulations.

Along with β-cells, the greater islet paracrine milieu is known to be affected in T2D. For example, oversecretion of glucagon from α-cells contributes to hyperglycemia. However, the precise mechanisms underlying glucagon secretion (and α-cell function more broadly) have been difficult to interrogate. Briant et al. (68) used β-cell–specific ChR2 expression to elucidate the paracrine regulation of glucagon secretion. Optogenetic activation of β-cells caused hyperpolarization of α-cells, with a marked reduction in action potential frequency and glucagon secretion. In contrast, β-cell activation led to rapid elevation of electrical activity in δ-cells, as well as increased secretion of SST. α-Cells express SST-2 receptors, and their pharmacological inhibition abolished the observed effects. It was concluded that SST secretion from δ-cells affects membrane potential and electrical activity of α-cells. In addition, electrical coupling via gap junctions was confirmed between β- and δ-cells. These findings point to a possible mechanism contributing to hyperglycemia in T2D, whereby islets develop reduced contacts between β- and δ-cells.

Paracrine regulation in the opposite direction, i.e., from δ- to β-cells, has been studied with the help of a δ-cell–specific ChR2 mouse model (SST-ChR2) (69). Islets were subsequently virally transduced with a β-cell–specific Ca2+ reporter and transplanted into the ACE of wild-type mice as an in situ model. The direct δ-cell activation led to a rapid reduction in β-cell Ca2+ oscillations as well as in the frequency and amplitude of their electrical activity (69). The work also demonstrated that δ-cells contain dynamic filopodia structures that comprise SST secretory machinery and allow a δ-cell to contact multiple β-cells. In a HFD prediabetic mouse model, δ-cell function and filopodia were altered, aligning with previous reports on altered SST responses in islets from diabetic animals (70,71).

Complementary to depolarization in endocrine cells, ChR2 has been applied to islet-innervating neurons. Neurons from the dorsal motor nucleus of the vagus nerve project to the pancreas and modulate glucose homeostasis. For example, insulin secretion has previously been observed to be regulated by electrical parasympathetic stimulation (72,73). However, the majority of parasympathetic axons innervate the abdominal and thoracic areas of the body, making it difficult to target pancreas-specific axons. Optogenetics has been used to overcome this limitation toward a more targeted understanding of islet-innervating neuronal function. As mentioned above, Yang et al. (66) showed that selective light-dependent inhibition of peri-islet neurons leads to decreased electrical coupling between β-cells and other endocrine cell types in zebrafish. Two further mechanistic studies have been performed using a transgenic mouse model comprising cholinergic cell-specific expression of the ChR2 gene under the control of the choline acetyltransferase promoter. Fontaine et al. (74) first demonstrated that light-dependent stimulation of vagus nerves in the pancreas using an optical cannula in anesthetized mice promotes glucose-stimulated insulin secretion as well as increase blood flow within the pancreatic vasculature. Kawana et al. (18) used a different illumination approach that relied on UCNPs in live animals. UCNPs were delivered via pancreatic intraductal injection, and animals were subsequently subjected to external NIR illumination for chronic stimulation (2–8 weeks) (Fig. 3C). Along with increased insulin secretion and blood flow, chronic stimulation promoted β-cell proliferation, β-cell mass expansion, and suppression of STZ-induced hyperglycemia. These results demonstrate the strong influence of vagal signals in regulating β-cell physiology and function.

Intracellular Signaling Dynamics

An emerging application of optogenetics is to dissect the inner workings of cells, in particular the wiring and dynamics of their signaling pathways. As mentioned above, membrane potential alterations are intricately linked to intracellular Ca2+ signals and, consequently, insulin release (Fig. 2). In addition, dynamic Ca2+ levels regulate genetic programs, in many cases through calcineurin and NFAT. Miranda et al. (75) used ChR2 to examine the links between β-cell electrical activity, Ca2+ oscillations, and NFAT-driven gene regulation. The authors first demonstrated that elevated glucose increased NFAT activity through a membrane potential– and Ca2+-dependent process. Temporal light patterns then induced Ca2+ bouts to dissect the signaling network. Notably, this analysis revealed that only relatively slow light patterns, akin to natural β-cell activity profiles, elicited NFAT activation. Thus β-cells, unlike other cell types (76), acted with the characteristics of a low-pass filter rather than a detector of cumulative Ca2+. The authors next investigated to what extent ChR2 activation induced transcriptional changes using transcriptomics. ChR2 stimulation resulted in up/downregulation of a smaller number of genes compared with glucose stimulation. These critical differences may reflect either a requirement for a different temporal NFAT activation pattern that was not recapitulated or effects of glucose through other pathways.

Signaling processes other than Ca2+ also play a role in β-cells. For instance, mitogenic and prosurvival activities are exerted through peptide growth factors and their cognate receptors, such as fibroblast growth factors acting on fibroblast growth factor receptors (FGFRs). Our group has used optogenetics to control the mitogen-activated protein kinases/extracellular signal–regulated kinase and phosphatidylinositol-3 kinase/Akt pathways downstream of growth factors (77,78). In INS-1E β-cells, the optogenetic method was a highly engineered fusion receptor consisting of a red light–sensitive phytochrome of cyanobacterial origin and the catalytic tyrosine kinase domain of murine FGFR1 (rOpto-FGFR1) (79). Due to the efficient tissue penetration of red light, and because of the exceptional light sensitivity of phytochromes, signaling could be activated even through several millimeters of synthetic tissue, comparable to the distance separating the pancreas from skin in mice. This proof-of-principle study demonstrated the potential of red light to manipulate activity in kinase-dependent signaling pathways deep in a mammalian tissue.

Notably, mechanistic signaling studies in islets have reached beyond endocrine cells. Pericytes are contractile mural cells that line capillaries and regulate vasculature stability and blood flow. The regulation of blood flow likely impacts the delivery of glucose to islets and in turn the release of glucoregulatory hormones into the circulation. Two recent studies used optogenetics to interrogate pericyte activity. In the first study, Tamayo et al. (80) expressed ChR2 in islet pericytes that were grafted into the retinas of STZ-treated diabetic mice. Pericyte activity was controlled using focused laser light (during in vivo imaging of islet blood flow) or whole-cage LED illumination (during long-term illumination of awake animals). Although only 50% of the identifiable pericyte population expressed ChR2, marked reversible reductions in capillary diameter and blood flow were observed. Importantly, these changes manifested in reduced islet glucose sensing and hormone secretion, leading ultimately to hyperglycemia. In a further study, Michau et al. (81) expressed ChR2 in pericytes in the intact, but exposed pancreas of anesthetized mice. Stimulation with blue light from an optical fiber also decreased blood flow, which was delayed in the case of diabetic animals that had been fed a HFD. Collectively, these two studies provided cell type–specific evidence on the involvement of pericytes in acute metabolic processes and, in turn, on how metabolic stress impacts pericyte function.

At the inception of the field in the mid-2000s, optogenetic methods have been applied in a handful of core laboratories. Ever since, several major waves of evolution have resulted in broader access to and uptake of optogenetics. First, the many available transgenic rodents and advanced viral systems designed to express actuators have been adapted to generate new, e.g., islet cell–specific models. Second, diverse light sources are now available to stimulate cells ex vivo and in vivo (Table 2 and Fig. 3). Ex vivo, these include dedicated LED illuminators, the most advanced of which even allow single-well control in multiwell plates (82,83), as well as widefield and confocal microscopes that may already be available in many laboratories. In vivo, these include the methods summarized above, the most advanced of which allow wireless, noninvasive light delivery (Table 2 and Fig. 3). Third, a wide range of cellular processes can now be controlled using light, including ionic signals, protein secretion, signal transduction, transcription, protein stability, and subcellular localization.

While impressive outcomes have been achieved as summarized in this review, it is worthwhile noting potential limitations of the technique. First, there always is the potential for cellular toxicity and undesired immune responses as a consequence of actuator delivery or expression (84,85). Second, expression of actuators can result in functional alterations in the targeted cells and tissues in the absence of light. The prominent causes of this dark activity are contaminating light from light sources other than the stimulation source or the intrinsic activity of some actuators in unstimulated states. Adequate controls and multiple independent assays can assist in limiting this (86,87). Third, while light signals can be controlled with high temporal precision (essentially limited by the performance of the light source and software), responses can act tardily. This is due to the residence time of some actuators in the light-activated state even after light has been withdrawn. Active state lifetimes can vary between seconds to many minutes or hours depending on the method applied (88–90). Fourth, some actuators are subject to peculiarities that need to be understood to collect reliable data toward meaningful conclusions. For instance, a paradoxical sporadic excitation of electrical activity can be observed in NpHR-expressing cells after extensive periods of light-induced inhibition (91,92). As a further example, some opsins suffer from decreased current activity during continuous stimulation or decreased responsiveness to repeated stimulation attempts (93,94). Furthermore, not all actuators are compatible with multiphoton excitation that provides highly confined deep tissue stimulation (95–97).

As an alternative to genetically encoded actuators, photopharmacology negates the need for gene delivery by using light-responsive drugs often constructed around photochromic moieties (e.g., azobenzenes). Photopharmacology has been used in studies of β-cell function and is particularly suitable for interrogation of endogenous receptor systems and signaling pathways (11,98). Photopharmacology calls for its own set of considerations and limitations, including a need to effectively deliver the photoactive drug and the undesired activation of all cell types that are exposed to the compound and express its target receptor.

Irrespective of the molecular nature of the modality, the use of light can be associated with undesirable side effects. Several studies have reported that light, in the absence of actuators, can induce cellular responses through either endogenously expressed opsins, heat, or photochemical effects (99–101). Such effects have been documented in other tissues and can be managed using controls (i.e., naive cells), red-shifted excitation light, dedicated cell culture media (102,103), and carefully engineered optical devices.

Pancreatic islets are complex and essential multicellular structures. In the past 5 years, primary functionalities of islet cells, their internal wiring, as well as their cell-cell interactions have been interrogated using optogenetics. Leaving from here, what are future perspectives for the field?

Optogenetic actuators have been harnessed from an array of nonmetazoan organisms, and the repertoire is impressive. For instance, the comprehensive OptoBase database (104) lists >40 light-sensitive core domains, many of which have been combined with signaling effectors. Remarkably, many actuator classes have not yet been used in islets, including light-activated transcription factors and genome modifiers (105), GPCRs other than OPN4 (106), or kinases (78,107). Consequently, many interesting and medically relevant processes are yet to be interrogated with optogenetics. Furthermore, development of novel actuators has become more achievable in recent years through new resources (108,109).

Optogenetics has now matured to the point where applications to humans are within reach. The retina has been the first target for optogenetic therapies, with currently several ongoing clinical trials (84,110,111). These all rely on the delivery of ChRs using adeno-associated viruses, and a recent first clinical report revealed some visual restoration in one patient with advanced retinitis pigmentosa (3). In light of this progress, it is important to consider possible translational barriers in the context of islet cells. For light-sensitive designer cells (e.g., replacement β-cells), challenges are well known from other cell replacement strategies. Most studies to date used HEK-293 cells or β-cell lines to optimize the light-sensitive method. However, due to immunogenicity, these and other cell types may require encapsulation toward further applications. Encapsulation is also important in protecting these cells from the host immune system and general environment while being permeable to oxygen and nutrients (112). More clinically applicable chassis, such as human mesenchymal stem cells (34) and induced pluripotent stem cells (49), are increasingly being pursued in genetic engineering and cell therapy. Other challenges include finding an appropriate implantation site, limiting fibrosis, improving vascularization methods using stable biocompatible materials that support cell survival and function, and ensuring that the materials are not immunogenic (113). A wide range of strategies are being explored; however, the most optimal for cellular diabetes therapy is yet to be established (112,113).

In addition to cell type selection and general cell therapy challenges, the optical hardware needs to fit the purpose (Table 2). At present, many studies in rodents have used external light sources (Fig. 3). More advanced coencapsulation of designer cells with LED implants has been achieved (44), as have closed-loop systems with continuous glucose sensing (31,33,45). However, optimal materials and conditions for safe and long-term optogenetic therapy are unknown. Further applications will likely require improved integration of hardware, cells, glucose monitoring, and encapsulation. Key considerations may involve reduction of heat; incorporation of lightweight, flexible, and stable materials that are light-permeable and limit an immune response from the hardware itself; and minimization of local tissue damage (114,115). Alternatively, light-transducing materials, such as UCNPs, that negate the need for implants, can also be coencapsulated with cells (45) and achieve light delivery deep into tissues (114). However, the particle concentration sufficient for visible light generation and actuator stimulation is critical, and the slow excitation response and long luminescence decay times of some UCNP systems may not be appropriate for controlling rapid processes. Finally, potential cytotoxicity of UCNPs in vivo needs to be assessed to determine clinical applicability (114). In general, direct stimulation of islets residing in the human pancreas using an illumination device will likely be very difficult because of its relatively large size and location in the peritoneal cavity.

While we have focused on islet cells and metabolic consequences of (dys)regulation, it is noteworthy that optogenetics has also been applied to other (patho)physiological processes related to the pancreas. For example, a recent study demonstrated the optical control of the nuclear factor-κB signaling pathway in a cancer cell line of pancreatic ductal origin to study disease development and progression (116). In combination with optical reporters, such as luciferases, light-sensitive models may also allow the identification of new active agents in optogenetics-assisted drug screens as demonstrated previously (117,118). These final examples demonstrate the breadth of opportunities that are offered by eliciting effects using light-sensitive actuators in pancreatic islets.

C.G.G. is currently affiliated with the Department of Developmental Biology, School of Medicine, Stanford University, Stanford, CA.

Acknowledgments. The authors thank their colleagues for critical advice on this manuscript.

Funding. This study was supported by a JDRF Australia PhD Top-up Scholarship (to C.G.G.), Australian Research Council grants FT200100519 and DP200102093 (to H.J.), and National Health and Medical Research Council grant APP1187638 (to H.J.). The Australian Regenerative Medicine Institute is supported by State Government of Victoria and the Australian Government grants. The European Molecular Biology Laboratory Australia Partnership Laboratory is supported by the National Collaborative Research Infrastructure Strategy of the Australian Government.

Duality of Interest. No potential conflicts of interest relevant to this article were reported.

Author Contributions. C.G.G. generated the figures and tables. C.G.G. and H.J. reviewed the current literature and contributed to the writing and editing of the manuscript. C.G.G. is the guarantor of this work and, as such, had full access to all the data in the study and takes responsibility for the integrity of the data and the accuracy of the data analysis.

Data and Resource Availability. No data sets and applicable resources were generated or analyzed for this article. All previously published data are referenced.

1.
Boyden
ES
,
Zhang
F
,
Bamberg
E
,
Nagel
G
,
Deisseroth
K
.
Millisecond-timescale, genetically targeted optical control of neural activity
.
Nat Neurosci
2005
;
8
:
1263
1268
2.
Zemelman
BV
,
Lee
GA
,
Ng
M
,
Miesenböck
G
.
Selective photostimulation of genetically chARGed neurons
.
Neuron
2002
;
33
:
15
22
3.
Sahel
J-A
,
Boulanger-Scemama
E
,
Pagot
C
, et al
.
Partial recovery of visual function in a blind patient after optogenetic therapy
.
Nat Med
2021
;
27
:
1223
1229
4.
Da Silva Xavier
G
.
The cells of the islets of Langerhans
.
J Clin Med
2018
;
7
:
54
5.
Rorsman
P
,
Ashcroft
FM
.
Pancreatic β-cell electrical activity and insulin secretion: of mice and men
.
Physiol Rev
2018
;
98
:
117
214
6.
Campbell
JE
,
Newgard
CB
.
Mechanisms controlling pancreatic islet cell function in insulin secretion
.
Nat Rev Mol Cell Biol
2021
;
22
:
142
158
7.
Rorsman
P
,
Huising
MO
.
The somatostatin-secreting pancreatic δ-cell in health and disease
.
Nat Rev Endocrinol
2018
;
14
:
404
414
8.
Prasad-Reddy
L
,
Isaacs
D
.
A clinical review of GLP-1 receptor agonists: efficacy and safety in diabetes and beyond
.
Drugs Context
2015
;
4
:
212283
9.
Nathan
DM
.
Long-term complications of diabetes mellitus
.
N Engl J Med
1993
;
328
:
1676
1685
10.
Chen
Z
,
Truskinovsky
L
,
Tzanakakis
ES
.
Emerging molecular technologies for light-mediated modulation of pancreatic beta-cell function
.
Mol Metab
2022
;
64
:
101552
11.
Frank
JA
,
Broichhagen
J
,
Yushchenko
DA
,
Trauner
D
,
Schultz
C
,
Hodson
DJ
.
Optical tools for understanding the complexity of β-cell signalling and insulin release
.
Nat Rev Endocrinol
2018
;
14
:
721
737
12.
Huey
J
,
Keutler
K
,
Schultz
C
.
Chemical Biology toolbox for studying pancreatic islet function - a perspective
.
Cell Chem Biol
2020
;
27
:
1015
1031
13.
Meister
J
,
Wang
L
,
Pydi
SP
,
Wess
J
.
Chemogenetic approaches to identify metabolically important GPCR signaling pathways: therapeutic implications
.
J Neurochem
2021
;
158
:
603
620
14.
Deisseroth
K
.
Optogenetics: 10 years of microbial opsins in neuroscience
.
Nat Neurosci
2015
;
18
:
1213
1225
15.
Kanno
T
,
Gopel
SO
,
Rorsman
P
,
Wakui
M
.
Cellular function in multicellular system for hormone-secretion: electrophysiological aspect of studies on alpha-, beta- and delta-cells of the pancreatic islet
.
Neurosci Res
2002
;
42
:
79
90
16.
Nagel
G
,
Szellas
T
,
Huhn
W
, et al
.
Channelrhodopsin-2, a directly light-gated cation-selective membrane channel
.
Proc Natl Acad Sci U S A
2003
;
100
:
13940
13945
17.
Zhang
F
,
Wang
L-P
,
Brauner
M
, et al
.
Multimodal fast optical interrogation of neural circuitry
.
Nature
2007
;
446
:
633
639
18.
Kawana
Y
,
Imai
J
,
Morizawa
YM
, et al
.
Optogenetic stimulation of vagal nerves for enhanced glucose-stimulated insulin secretion and β cell proliferation
.
Nat Biomed Eng
2024
;
8
:
808
822
19.
Armbruster
BN
,
Li
X
,
Pausch
MH
,
Herlitze
S
,
Roth
BL
.
Evolving the lock to fit the key to create a family of G protein-coupled receptors potently activated by an inert ligand
.
Proc Natl Acad Sci U S A
2007
;
104
:
5163
5168
20.
Sternson
SM
,
Roth
BL
.
Chemogenetic tools to interrogate brain functions
.
Annu Rev Neurosci
2014
;
37
:
387
407
21.
Wess
J
.
In vivo metabolic roles of G proteins of the Gi family studied with novel mouse models
.
Endocrinology
2022
;
163
:
bqab245
22.
Speier
S
,
Nyqvist
D
,
Köhler
M
,
Caicedo
A
,
Leibiger
IB
,
Berggren
P-O
.
Noninvasive high-resolution in vivo imaging of cell biology in the anterior chamber of the mouse eye
.
Nat Protoc
2008
;
3
:
1278
1286
23.
Yang
S-N
,
Shi
Y
,
Berggren
P-O
.
The anterior chamber of the eye technology and its anatomical, optical, and immunological bases
.
Physiol Rev
2024
;
104
:
881
929
24.
Hososhima
S
,
Yuasa
H
,
Ishizuka
T
, et al
.
Near-infrared (NIR) up-conversion optogenetics
.
Sci Rep
2015
;
5
:
16533
25.
All
AH
,
Zeng
X
,
Teh
DBL
, et al
.
Expanding the toolbox of upconversion nanoparticles for in vivo optogenetics and neuromodulation
.
Adv Mater
2019
;
31
:
e1803474
26.
Hedeskov
CJ
.
Mechanism of glucose-induced insulin secretion
.
Physiol Rev
1980
;
60
:
442
509
27.
Wang
X
,
Chen
X
,
Yang
Y
.
Spatiotemporal control of gene expression by a light-switchable transgene system
.
Nat Methods
2012
;
9
:
266
269
28.
Li
T
,
Chen
X
,
Qian
Y
, et al
.
A synthetic BRET-based optogenetic device for pulsatile transgene expression enabling glucose homeostasis in mice
.
Nature Commun
2021
;
12
:
1
10
29.
Li
C-Y
,
Wu
T
,
Zhao
X-J
, et al
.
A glucose-blue light AND gate-controlled chemi-optogenetic cell-implanted therapy for treating type-1 diabetes in mice
.
Front Bioeng Biotechnol
2023
;
11
:
1052607
30.
Yazawa
M
,
Sadaghiani
AM
,
Hsueh
B
,
Dolmetsch
RE
.
Induction of protein-protein interactions in live cells using light
.
Nat Biotechnol
2009
;
27
:
941
945
31.
Shao
J
,
Xue
S
,
Yu
G
, et al
.
Smartphone-controlled optogenetically engineered cells enable semiautomatic glucose homeostasis in diabetic mice
.
Sci Transl Med
2017
;
9
:
eaal2298
32.
Yu
G
,
Yu
Y
,
Ye
H
. Constructing a smartphone-controlled semiautomatic theranostic system for glucose homeostasis in diabetic mice.
Methods Mol Biol
2021
;
2312
:
141
-
158
33.
Man
T
,
Yu
G
,
Zhu
F
, et al
.
Antidiabetic close loop based on wearable DNA-hydrogel glucometer and implantable optogenetic cells
.
JACS Au
2024
;
4
:
1500
1508
34.
Yu
G
,
Zhang
M
,
Gao
L
, et al
.
Far-red light-activated human islet-like designer cells enable sustained fine-tuned secretion of insulin for glucose control
.
Mol Ther
2022
;
30
:
341
354
35.
Zhou
Y
,
Kong
D
,
Wang
X
, et al
.
A small and highly sensitive red/far-red optogenetic switch for applications in mammals
.
Nat Biotechnol
2022
;
40
:
262
272
36.
Possart
A
,
Hiltbrunner
A
.
An evolutionarily conserved signaling mechanism mediates far-red light responses in land plants
.
Plant Cell
2013
;
25
:
102
114
37.
Rausenberger
J
,
Tscheuschler
A
,
Nordmeier
W
, et al
.
Photoconversion and nuclear trafficking cycles determine phytochrome A’s response profile to far-red light
.
Cell
2011
;
146
:
813
825
38.
Drucker
DJ
,
Philippe
J
,
Mojsov
S
,
Chick
WL
,
Habener
JF
.
Glucagon-like peptide I stimulates insulin gene expression and increases cyclic AMP levels in a rat islet cell line
.
Proc Natl Acad Sci U S A
1987
;
84
:
3434
3438
39.
Ye
H
,
Daoud-El Baba
M
,
Peng
R-W
,
Fussenegger
M
.
A synthetic optogenetic transcription device enhances blood-glucose homeostasis in mice
.
Science
2011
;
332
:
1565
1568
40.
Choi
M
,
Choi
JW
,
Kim
S
,
Nizamoglu
S
,
Hahn
SK
,
Yun
SH
.
Light-guiding hydrogels for cell-based sensing and optogenetic synthesis in vivo
.
Nat Photonics
2013
;
7
:
987
994
41.
Jost
M
,
Fernández-Zapata
J
,
Polanco
MC
, et al
.
Structural basis for gene regulation by a B12-dependent photoreceptor
.
Nature
2015
;
526
:
536
541
42.
Kainrath
S
,
Stadler
M
,
Reichhart
E
,
Distel
M
,
Janovjak
H
.
Green-light-induced inactivation of receptor signaling using cobalamin-binding domains
.
Angew Chem Int Ed Engl
2017
;
56
:
4608
4611
43.
Mansouri
M
,
Hussherr
M-D
,
Strittmatter
T
, et al
.
Smart-watch-programmed green-light-operated percutaneous control of therapeutic transgenes
.
Nat Commun
2021
;
12
:
3388
44.
Liu
Z
,
Zhou
Y
,
Qu
X
, et al
.
A self-powered optogenetic system for implantable blood glucose control
.
Research (Wash D C)
2022
;
2022
:
9864734
45.
Lu
Q
,
Wang
Z
,
Bai
S
, et al
.
Hydrophobicity regulation of energy acceptors confined in mesoporous silica enabled reversible activation of optogenetics for closed-loop glycemic control
.
J Am Chem Soc
2023
;
145
:
5941
5951
46.
Mansouri
M
,
Xue
S
,
Hussherr
M-D
,
Strittmatter
T
,
Camenisch
G
,
Fussenegger
M
.
Smartphone-flashlight-mediated remote control of rapid insulin secretion restores glucose homeostasis in experimental type-1 diabetes
.
Small
2021
;
17
:
e2101939
47.
Reinbothe
TM
,
Safi
F
,
Axelsson
AS
,
Mollet
IG
,
Rosengren
AH
.
Optogenetic control of insulin secretion in intact pancreatic islets with β-cell-specific expression of channelrhodopsin-2
.
Islets
2014
;
6
:
e28095
48.
Kushibiki
T
,
Okawa
S
,
Hirasawa
T
,
Ishihara
M
.
Optogenetic control of insulin secretion by pancreatic β-cells in vitro and in vivo
.
Gene Ther
2015
;
22
:
553
559
49.
Choi
J
,
Shin
E
,
Lee
J
, et al
.
Light-stimulated insulin secretion from pancreatic islet-like organoids derived from human pluripotent stem cells
.
Mol Ther
2023
;
31
:
1480
1495
50.
Kim
S
,
Kyung
T
,
Chung
J-H
, et al
.
Non-invasive optical control of endogenous Ca2+ channels in awake mice
.
Nature Commun
2020
;
11
:
210
51.
Kennedy
MJ
,
Hughes
RM
,
Peteya
LA
,
Schwartz
JW
,
Ehlers
MD
,
Tucker
CL
.
Rapid blue-light-mediated induction of protein interactions in living cells
.
Nat Methods
2010
;
7
:
973
975
52.
Kyung
T
,
Lee
S
,
Kim
JE
, et al
.
Optogenetic control of endogenous Ca(2+) channels in vivo
.
Nat Biotechnol
2015
;
33
:
1092
1096
53.
Zhang
F
,
Tzanakakis
ES
.
Optogenetic regulation of insulin secretion in pancreatic β-cells
.
Sci Rep
2017
;
7
:
9357
54.
Zhang
F
,
Tzanakakis
ES
.
Amelioration of diabetes in a murine model upon transplantation of pancreatic β-cells with optogenetic control of cyclic adenosine monophosphate
.
ACS Synth Biol
2019
;
8
:
2248
2255
55.
Chen
Z
,
Stoukides
DM
,
Tzanakakis
ES
.
Light-mediated enhancement of glucose-stimulated insulin release of optogenetically engineered human pancreatic beta-cells
.
ACS Synth Biol
2024
;
13
:
825
836
56.
Mansouri
M
,
Ray
PG
,
Franko
N
,
Xue
S
,
Fussenegger
M
.
Design of programmable post-translational switch control platform for on-demand protein secretion in mammalian cells
.
Nucleic Acids Res
2023
;
51
:
e1
57.
Wang
X
,
Kang
L
,
Kong
D
, et al
.
A programmable protease-based protein secretion platform for therapeutic applications
.
Nat Chem Biol
2023
;
20
:
432
442
58.
Kawano
F
,
Suzuki
H
,
Furuya
A
,
Sato
M
.
Engineered pairs of distinct photoswitches for optogenetic control of cellular proteins
.
Nat Commun
2015
;
6
:
6256
59.
Jain
R
,
Lammert
E
.
Cell–cell interactions in the endocrine pancreas
.
Diabetes Obes Metab
2009
;
11
(
Suppl. 4
):
159
167
60.
Benninger
RKP
,
Kravets
V
.
The physiological role of β-cell heterogeneity in pancreatic islet function
.
Nat Rev Endocrinol
2022
;
18
:
9
22
61.
Johnston
NR
,
Mitchell
RK
,
Haythorne
E
, et al
.
Beta cell hubs dictate pancreatic islet responses to glucose
.
Cell Metab
2016
;
24
:
389
401
62.
El-Gaby
M
,
Zhang
Y
,
Wolf
K
,
Schwiening
CJ
,
Paulsen
O
,
Shipton
OA
.
Archaerhodopsin selectively and reversibly silences synaptic transmission through altered pH
.
Cell Rep
2016
;
16
:
2259
2268
63.
Mahn
M
,
Gibor
L
,
Patil
P
, et al
.
High-efficiency optogenetic silencing with soma-targeted anion-conducting channelrhodopsins
.
Nat Commun
2018
;
9
:
4125
64.
Wietek
J
,
Wiegert
JS
,
Adeishvili
N
, et al
.
Conversion of channelrhodopsin into a light-gated chloride channel
.
Science
2014
;
344
:
409
412
65.
Souslova
EA
,
Mironova
KE
,
Deyev
SM
.
Applications of genetically encoded photosensitizer miniSOG: from correlative light electron microscopy to immunophotosensitizing
.
J Biophotonics
2017
;
10
:
338
352
66.
Yang
YHC
,
Briant
LJB
,
Raab
CA
, et al
.
Innervation modulates the functional connectivity between pancreatic endocrine cells
.
Elife
2022
;
11
:
e64526
67.
Westacott
MJ
,
Ludin
NWF
,
Benninger
RKP
.
Spatially organized β-cell subpopulations control electrical dynamics across islets of Langerhans
.
Biophys J
2017
;
113
:
1093
1108
68.
Briant
LJB
,
Reinbothe
TM
,
Spiliotis
I
,
Miranda
C
,
Rodriguez
B
,
Rorsman
P
.
δ-Cells and β-cells are electrically coupled and regulate α-cell activity via somatostatin
.
J Physiol
2018
;
596
:
197
215
69.
Arrojo E Drigo
R
,
Jacob
S
,
García-Prieto
CF
, et al
.
Structural basis for delta cell paracrine regulation in pancreatic islets
.
Nat Commun
2019
;
10
:
3700
3712
70.
Abdel-Halim
SM
,
Guenifi
A
,
Efendić
S
,
Ostenson
CG
.
Both somatostatin and insulin responses to glucose are impaired in the perfused pancreas of the spontaneously noninsulin-dependent diabetic GK (Goto-Kakizaki) rats
.
Acta Physiol Scand
1993
;
148
:
219
226
71.
Collins
SC
,
Salehi
A
,
Eliasson
L
,
Olofsson
CS
,
Rorsman
P
.
Long-term exposure of mouse pancreatic islets to oleate or palmitate results in reduced glucose-induced somatostatin and oversecretion of glucagon
.
Diabetologia
2008
;
51
:
1689
1693
72.
Frohman
LA
,
Ezdinli
EZ
,
Javid
R
.
Effect of vagotomy and vagal stimulation on insulin secretion
.
Diabetes
1967
;
16
:
443
448
73.
Daniel
PM
,
Henderson
JR
.
The effect of vagal stimulation on plasma insulin and glucose levels in the baboon
.
J Physiol
1967
;
192
:
317
327
74.
Fontaine
AK
,
Ramirez
DG
,
Littich
SF
, et al
.
Optogenetic stimulation of cholinergic fibers for the modulation of insulin and glycemia
.
Sci Rep
2021
;
11
:
3670
3679
75.
Miranda
JG
,
Schleicher
WE
,
Wells
KL
,
Ramirez
DG
,
Landgrave
SP
,
Benninger
RKP
.
Dynamic changes in β-cell [Ca2+] regulate NFAT activation, gene transcription, and islet gap junction communication
.
Mol Metab
2022
;
57
:
101430
76.
Hannanta-Anan
P
,
Chow
BY
.
Optogenetic control of calcium oscillation waveform defines NFAT as an integrator of calcium load
.
Cell Syst
2016
;
2
:
283
288
77.
Grusch
M
,
Schelch
K
,
Riedler
R
, et al
.
Spatio-temporally precise activation of engineered receptor tyrosine kinases by light
.
Embo J
2014
;
33
:
1713
1726
78.
Crossman
SH
,
Janovjak
H
.
Light-activated receptor tyrosine kinases: designs and applications
.
Curr Opin Pharmacol
2022
;
63
:
102197
79.
Reichhart
E
,
Ingles-Prieto
A
,
Tichy
A-M
,
McKenzie
C
,
Janovjak
H
.
A phytochrome sensory domain permits receptor activation by red light
.
Angew Chem Int Ed Engl
2016
;
55
:
6339
6342
80.
Tamayo
A
,
Gonçalves
LM
,
Rodriguez-Diaz
R
, et al
.
Pericyte control of blood flow in intraocular islet grafts impacts glucose homeostasis in mice
.
Diabetes
2022
;
71
:
1679
1693
81.
Michau
A
,
Lafont
C
,
Bargi-Souza
P
, et al
.
Metabolic stress impairs pericyte response to optogenetic stimulation in pancreatic islets
.
Front Endocrinol (Lausanne)
2022
;
13
:
918733
82.
Bugaj
LJ
,
Lim
WA
.
High-throughput multicolor optogenetics in microwell plates
.
Nat Protoc
2019
;
14
:
2205
2228
83.
Gangemi
CG
,
Janovjak
H
.
Modular light-emitting diode shelving systems for scalable optogenetics
. 12 April
2023
[preprint].
Research Square
:
PPR644592
84.
Shen
Y
,
Campbell
RE
,
Côté
DC
,
Paquet
M-E
.
Challenges for therapeutic applications of opsin-based optogenetic tools in humans
.
Front Neural Circuits
2020
;
14
:
41
85.
Maimon
BE
,
Diaz
M
,
Revol
ECM
, et al
.
Optogenetic peripheral nerve immunogenicity
.
Sci Rep
2018
;
8
:
14076
86.
Kainrath
S
,
Janovjak
H
. Design and application of light-regulated receptor tyrosine kinases. In
Photoswitching Proteins: Methods and Protocols.
Niopek
D
, Ed.
New York
,
Springer US
,
2020
, pp.
233
246
87.
Khamo
JS
,
Krishnamurthy
VV
,
Chen
Q
,
Diao
J
,
Zhang
K
.
Optogenetic delineation of receptor tyrosine kinase subcircuits in PC12 cell differentiation
.
Cell Chem Biol
2019
;
26
:
400
410.
e3
88.
Spoida
K
,
Eickelbeck
D
,
Karapinar
R
, et al
.
Melanopsin variants as intrinsic optogenetic on and off switches for transient versus sustained activation of G protein pathways
.
Curr Biol
2016
;
26
:
1206
1212
89.
Berndt
A
,
Yizhar
O
,
Gunaydin
LA
,
Hegemann
P
,
Deisseroth
K
.
Bi-stable neural state switches
.
Nat Neurosci
2009
;
12
:
229
234
90.
Benedetti
L
,
Barentine
AES
,
Messa
M
,
Wheeler
H
,
Bewersdorf
J
,
De Camilli
P
.
Light-activated protein interaction with high spatial subcellular confinement
.
Proc Natl Acad Sci U S A
2018
;
115
:
e2238
e2245
91.
Arrenberg
AB
,
Del Bene
F
,
Baier
H
.
Optical control of zebrafish behavior with halorhodopsin
.
Proc Natl Acad Sci U S A
2009
;
106
:
17968
17973
92.
Raimondo
JV
,
Kay
L
,
Ellender
TJ
,
Akerman
CJ
.
Optogenetic silencing strategies differ in their effects on inhibitory synaptic transmission
.
Nat Neurosci
2012
;
15
:
1102
1104
93.
Lin
JY
.
A user’s guide to channelrhodopsin variants: features, limitations and future developments
.
Exp Physiol
2011
;
96
:
19
25
94.
Schoenenberger
P
,
Schärer
Y-PZ
,
Oertner
TG
.
Channelrhodopsin as a tool to investigate synaptic transmission and plasticity
.
Exp Physiol
2011
;
96
:
34
39
95.
Oron
D
,
Papagiakoumou
E
,
Anselmi
F
,
Emiliani
V
.
Two-photon optogenetics
.
Prog Brain Res
2012
;
196
:
119
143
96.
De La Crompe
B
,
Coulon
P
,
Diester
I
.
Functional interrogation of neural circuits with virally transmitted optogenetic tools
.
J Neurosci Methods
2020
;
345
:
108905
97.
Prakash
R
,
Yizhar
O
,
Grewe
B
, et al
.
Two-photon optogenetic toolbox for fast inhibition, excitation and bistable modulation
.
Nat Methods
2012
;
9
:
1171
1179
98.
Nasteska
D
,
Hodson
DJ
.
The role of beta cell heterogeneity in islet function and insulin release
.
J Mol Endocrinol
2018
;
61
:
R43
R60
99.
Grzelak
A
,
Rychlik
B
,
Bartosz
G
.
Light-dependent generation of reactive oxygen species in cell culture media
.
Free Radic Biol Med
2001
;
30
:
1418
1425
100.
Icha
J
,
Weber
M
,
Waters
JC
,
Norden
C
.
Phototoxicity in live fluorescence microscopy, and how to avoid it
.
Bioessays
2017
;
39
:
1700003
101.
Tyssowski
KM
,
Gray
JM
.
Blue light increases neuronal activity-regulated gene expression in the absence of optogenetic proteins
.
ENeuro
2019
;
6
:ENEURO.0085-19.2019
102.
Stockley
JH
,
Evans
K
,
Matthey
M
, et al
.
Surpassing light-induced cell damage in vitro with novel cell culture media
.
Sci Rep
2017
;
7
:
849
103.
Zabolocki
M
,
McCormack
K
,
van den Hurk
M
, et al
.
BrainPhys neuronal medium optimized for imaging and optogenetics in vitro
.
Nat Commun
2020
;
11
:
5550
104.
Kolar
K
,
Knobloch
C
,
Stork
H
,
Žnidarič
M
,
Weber
W
.
OptoBase: a web platform for molecular optogenetics
.
ACS Synth Biol
2018
;
7
:
1825
1828
105.
Lan
T-H
,
He
L
,
Huang
Y
,
Zhou
Y
.
Optogenetics for transcriptional programming and genetic engineering
.
Trends Genet
2022
;
38
:
1253
1270
106.
Tichy
A-M
,
Gerrard
EJ
,
Sexton
PM
,
Janovjak
H
.
Light-activated chimeric GPCRs: limitations and opportunities
.
Curr Opin Struct Biol
2019
;
57
:
196
203
107.
Beyer
HM
,
Naumann
S
,
Weber
W
,
Radziwill
G
.
Optogenetic control of signaling in mammalian cells
.
Biotechnol J
2015
;
10
:
273
283
108.
Tichy
A-M
,
Gerrard
EJ
,
Legrand
JMD
,
Hobbs
RM
,
Janovjak
H
.
Engineering strategy and vector library for the rapid generation of modular light-controlled protein–protein interactions
.
J Mol Biol
2019
;
431
:
3046
3055
109.
Dagliyan
O
,
Dokholyan
NV
,
Hahn
KM
.
Engineering proteins for allosteric control by light or ligands
.
Nat Protoc
2019
;
14
:
1863
1883
110.
Janovjak
H
,
Kleinlogel
S
.
Optogenetic neuroregeneration
.
Neural Regen Res
2022
;
17
:
1468
1470
111.
White
M
,
Mackay
M
,
Whittaker
R
.
Taking optogenetics into the human brain: opportunities and challenges in clinical trial design
.
Open Access J Clin Trials
2020
;
2020
:
33
41
112.
Facklam
AL
,
Volpatti
LR
,
Anderson
DG
.
Biomaterials for personalized cell therapy
.
Adv Mater
2020
;
32
:
e1902005
113.
Farina
M
,
Alexander
JF
,
Thekkedath
U
,
Ferrari
M
,
Grattoni
A
.
Cell encapsulation: overcoming barriers in cell transplantation in diabetes and beyond
.
Adv Drug Deliv Rev
2019
;
139
:
92
115
114.
Bansal
A
,
Shikha
S
,
Zhang
Y
.
Towards translational optogenetics
.
Nat Biomed Eng
2023
;
7
:
349
369
115.
Guan
N
,
Gao
X
,
Ye
H
.
Engineering of optogenetic devices for biomedical applications in mammalian synthetic biology
.
Eng Biol
2022
;
6
:
35
49
116.
Stierschneider
A
,
Grünstäudl
P
,
Colleselli
K
, et al
.
Light-inducible spatio-temporal control of TLR4 and NF-κB-Gluc reporter in human pancreatic cell line
.
Int J Mol Sci
2021
;
22
:
9232
117.
Dempsey
GT
,
Chaudhary
KW
,
Atwater
N
, et al
.
Cardiotoxicity screening with simultaneous optogenetic pacing, voltage imaging and calcium imaging
.
J Pharmacol Toxicol Methods
2016
;
81
:
240
250
118.
Agus
V
,
Janovjak
H
.
Optogenetic methods in drug screening: technologies and applications
.
Curr Opin Biotechnol
2017
;
48
:
8
14
Readers may use this article as long as the work is properly cited, the use is educational and not for profit, and the work is not altered. More information is available at https://www.diabetesjournals.org/journals/pages/license.