Open access peer-reviewed chapter - ONLINE FIRST

Retinal Mechanical Sensation

Written By

Ji-Jie Pang

Submitted: 21 February 2024 Reviewed: 03 April 2024 Published: 06 May 2024

DOI: 10.5772/intechopen.114957

Cell Communication and Signaling in Health and Disease IntechOpen
Cell Communication and Signaling in Health and Disease Edited by Thomas Heinbockel

From the Edited Volume

Cell Communication and Signaling in Health and Disease [Working Title]

Dr. Thomas Heinbockel

Chapter metrics overview

13 Chapter Downloads

View Full Metrics

Abstract

Retinal neurons process light signals and respond to mechanical signals. mechanosensitive channels (MSCs) have been revealed in all retinal layers in humans, monkeys, mice, rats, porcine, salamanders, goldfish, etc. Some MSCs open in physiological conditions to regulate membrane potential, light responses, and neurotransmitter release, and some MSCs can mediate neurodegenerative effects. Alterations in the intraocular and external pressure critically involve the pathogenesis of glaucoma, traumatic retinal injury (TRI), and other retinal disorders. Our team revealed several MSCs in the outer and inner retinal neurons and first reported the pressure-evoked current and voltage response in salamander photoreceptors and primate bipolar cells. It is still unclear how retinal light pathways deal with endogenous and exogenous mechanical stimulation, and the physiological and pathological significance for retinal neurons to express multiple types of MSCs is not fully understood. This chapter will focus on the variety and functions of MSCs permeable to K+, Na+, and Ca2+, primarily including the big potassium channel (BK), two-pore domain potassium channel TRAAK and TREK, Piezo, epithelial sodium channel (ENaC), transient receptor potential channel vanilloid (TRPV) TRPV1, TRPV2, TRPV4, etc., in retinal photoreceptors, bipolar cells, horizontal cells, amacrine cells, and ganglion cells.

Keywords

  • retina
  • mechanosensitive channel (MSC)
  • glaucoma
  • traumatic retinal injury
  • light pathway
  • photoreceptor
  • bipolar cell
  • Amacrine cell
  • horizontal cell
  • retinal ganglion cell
  • mammal
  • vertebrate
  • patch-clamp
  • synaptic current
  • channel current
  • pharmacology
  • immunocytochemistry
  • PCR
  • genetic mutation
  • BK
  • TRAAK
  • TREK
  • TRPV
  • ENaC
  • piezo
  • TRP

1. Introduction

Pressure stresses are associated with multiple retinal disorders [1], such as acute and chronic glaucoma, traumatic retinal injury (TRI), and retinal dysfunctions that occur during air travel, diving, and mountain hiking. Glaucoma is a blinding disease characterized by the elevation of the intraocular pressure (IOP) and histological retinal damage primarily observed in the optic disk and retinal ganglion cells (RGCs). The worldwide prevalence of glaucoma in the population above 40 years is ∼3·5% [2, 3]. Glaucoma patients typically develop tunnel vision, a severe loss of the peripheral and paracentral visual fields. The higher vulnerability of RGCs and the peripheral vision to mechanical stresses are generally similar between glaucoma and the pressurized air wave-caused TRI [3, 4, 5, 6, 7].

In mammals like humans [8, 9, 10] and mice [11, 12], cone photoreceptors have a lower light sensitivity and a distribution peak at the central retina/fovea, which provides the best visual resolution for the central retina during daytime (except for S-cone that has a gradually higher density in the ventral retina [13, 14, 15]). Rods, on the other hand, dominantly occupy vast regions of the peripheral and paracentral retina [16], offering higher light sensitivity. Rod signals go through three pathways to reach RGCs, (1) the primary rod pathway formed by Rods➔Rod bipolar cells (RBCs)➔AII amacrine cells (AII ACs)---Cone-driven depolarizing BCs (cDBCs)➔RGCs, (2) the secondary rod pathway formed by Rods---Cones ➔ cBCs➔RGCs, and (3) the tertiary rod pathway composed by Rods➔ Hyperpolarizing BCs (HBCs)➔RGCs, where “➔” indicates glutamatergic synapses and “---” indicates electric synapses. The central vision is dominated by cones, and cone signals may reach RGCs via Cones➔cDBCs➔ON RGCs and ON-OFF RGCs and Cones➔cHBCs➔OFF RGCs and ON-OFF RGCs. It is uncertain whether mechanical signals can differentially affect retinal neurons in the rod and cone pathways.

Do retinal neurons express mechanosensitive ion channels (MSCs)? The answer is yes. All eukaryotes express MSCs to cope with physical stimuli [17]. In multiple mammalian species, MSCs have been observed in neurons and processes in the outer segment layer (OSL), inner segment layer (ISL), outer nuclear layer (ONL), outer plexiform layer (OPL), inner nuclear layer (INL, including somas of BCs and ACs), inner plexiform layer (IPL), ganglion cell layer (GCL, including somas of RGCs and displaced ACs), and nerve fiber layer (NFL), such as the big potassium channel (BK), two-pore domain potassium channel TRAAK and TREK, Piezo, epithelial sodium channel (ENaC), and transient receptor potential channel vanilloid TRPV1, TRPV2, and TRPV4. These MSCs have been confirmed as mechano-gated channels in sensory neurons, Xenopus oocytes, or cultured cell lines [18, 19], as they open directly by membrane tension and transduce the mechanical signals into electrical currents. Retinal neurons show variable levels and distribution patterns of MSCs, implying cell-type specific mechanical-dealing strategies and pressure effects.

MSCs may be activated by membrane stretch, mechanical and osmotic pressures, changes in osmolarity and temperature, membrane stretches, pH, and other modulators. The physical factors are often present in the ocular environment in physiological conditions, and some channels are known to be constitutionally open (see below). The retina lines the inner surface of the posterior eyeball, and a positive IOP is homogeneously maintained inside the eyeball by the volume of the aqueous humor against the limited space formed by the relatively rigid eyeball shell. Although the circulation of the aqueous humor critically regulates the IOP level, IOP still fluctuates. The eyeball shell is primarily formed by the sclera and cornea, and with a bulk modulus of ∼2.2 × 109 Pa, water is incompressible under normal conditions. Thus, any increase in aqueous humor volume would expand the eyeball space to stretch the retina. The elevated IOP in glaucoma can significantly enlarge eyeballs in children, known as buphthalmos, and it also focally expands the retina at the cupping region of the optic disk in chronic glaucoma patients. Besides, aging may directly induce structural remodeling and retinal deformation [12]. Therefore, MSCs in retinal neurons may be activated in physiological and pathological conditions. This chapter will briefly review the literature on several representative MSCs in retinal photoreceptors, BCs, horizontal cells (HCs), ACs, and RGCs for a better understanding of the significance of MSCs in retinal light pathways.

Advertisement

2. Ocular mechanical hemostasis

Stress and strain, like IOP and retinal stretch, are a pair of inseparable physical parameters that can activate MSCs. The eyeball structure resembles a hollow spherical shell with an enclosed space and a positive inner pressure. Normal IOP in human eyeballs is 10–21 mmHg (1.3–2.8 kPa) higher than the atmospheric pressure, and the pressure source is aqueous humor. Aqueous humor flows from the ciliary body to the posterior and anterior chambers and drains into the canal of Schlemm and uveoscleral vein. Aqueous humor contains 99.9% of water [20] and molecules like amino acids, proteins, and others like those found in blood plasma [21, 22]. Its formation involves three mechanisms [21, 23, 24]: (a) the ultrafiltration, that is, water and water-soluble substances pass the fenestrated ciliary capillary endothelium and enter into the ciliary interstitium along the hydrostatic and osmotic pressure gradients, (b) diffusion of water- and lipid-soluble substances from ciliary processes to the posterior chamber, and (c) active transportation/secretion of anions, cations, and other molecules by the non-pigmented epithelial cells at ciliary processes. A majority (80–90%) of aqueous humor is produced by active transportation and is not pressure-dependent.

Glaucoma is generally classified as high-tension glaucoma and normal-tension glaucoma, and both are diagnosable by similar pathological changes at the optic disc and the loss of visual field. High-tension glaucoma is characterized by the accumulation of aqueous humor due to the blockage of the iris or trabecular meshwork on the drainage of the aqueous humor at the area of the anterior chamber angle, including angle closure glaucoma and open-angle glaucoma.

IOP level is primarily determined by the volume of aqueous humor and has been expressed by PiIOP=Pe+FinFu/Ctrab [23, 24], where Pe is the episcleral venous pressure, Fin is the inflow/formation rate, Fu is the outflow rate via uveoscleral pathway, and Ctrab is the facility of the outflow through the trabecular pathway. The equation, however, does not include the ocular elasticity and volumes as variables affecting IOP.

The flow rate of the aqueous humor and IOP both fluctuate. The circulation of the aqueous humor acts to stabilize IOP, but IOP still changes with the heartbeat, breath, and exercise in a range of 2–10 mmHg (reviewed by Ref. [1]). Young’s modulus (E, tensile elasticity) was measured as 1.5 × 105 to 8.3 × 105 Pa for the sclera and 2.5 × 104 to 2.4 × 105 Pa in the cornea, and water has a bulk modulus (K) of 2.2 × 109 Pa and is incompressible under normal conditions. Recently, we reported the first mathematical equation with the modified bulk (K) modulus and Young’s modulus (E) for the relationship between ocular elasticity, aqueous humor volume, eyeball volume, the thickness of the eyeball shell, and IOP [25]. The equation (eq) is presented in the following six forms:

ΔIOP=KsVWVsE1
ΔIOP=1.33HREsVWVsE2
Vs=KsVWΔIOPE3
Vs=1.33HREsVWΔIOPE4
VWVs=ΔIOPKsE5
VWVs=ΔIOP1.33EsRHE6

where VS is the eyeball volume, VW is the volume of aqueous humor, IOP is the intraocular pressure, Δ indicates changes in the parameters, KS is the bulk modulus of the eyeball, ES is Young’s modulus of the eyeball shell, H is the thickness of the eyeball shell, and R is the inner radius of the eyeball. Assuming the aqueous humor is fully regulated, Eqs. (1) and (2) state that IOP alteration is positively correlated with the elasticity and the volume fluctuation of the aqueous humor relative to the eyeball volume. Eqs. (3) and (4) state that the eyeball volume is positively correlated with the elasticity and the volume fluctuation of the aqueous humor relative to the IOP change. A larger ΔVW/ΔIOP (severe accumulation of aqueous humor with less IOP elevation) predicts more intensive eyeball and retinal expansion. Eqs. (5) and (6) state that the eyeball expansion rate is negatively correlated with the eyeball elasticity and positively correlated with the IOP elevation.

Advertisement

3. Physiological properties of MSCs and MSCs expressed in photoreceptors

3.1 MSCs permeable to K+

BK channel is also known as the calcium- and voltage-gated large conductance potassium channel, Maxi-K, KCNMA1, Slo1, Kca1.1, and stretch-activated potassium channel. It is a homotetrameric channel of four identical, pore-forming, and Ca2+ and voltage-sensing units with or without associating with the regulatory β and γ subunits. The single-channel conductance is between 200 and 300 pS (reviewed by Ref. [26]).

BKs were reported in rods in goldfish [27] and salamander retinas [28, 29, 30]. They were expressed in the axon terminal [27], soma, and inner segment (IS) [28, 31] of rods and associated with membrane hyperpolarization and glutamate release [29]. BK at the presynaptic terminals of goldfish bipolar cells [27] mediates calcium-activated potassium current (I(KCa)). BK also influences calcium-mediated glutamate release in salamander rods [29] under normal conditions and provides an effective “clamp” of the dark membrane potential [31] to its normal resting level. The I(KCa) was largely inhibited by 100 nM iberiotoxin, a BK blocker, and sensitive to tetraethylammonium (TEA) [29]. BK blocker iberiotoxin increases the light response of some rods [30]. BK may be activated by flupirtine [32]. Cones from the rabbit and salamander retina also express BK [30, 33]. Studies from other tissues have revealed the mechanical sensitivity of BKs [34, 35, 36].

Mechanosensitive P2K channels (TRAAK and TREK) [37, 38, 39, 40, 41, 42] are gated primarily by membrane tension, and their opening is facilitated by convex membrane deformation [43, 44, 45]. Activation of the channel by mechanical stimulus is graded and reversible. The negative pressure that can activate half-maximally TREK-1 and TRAAK is −36 and −46 mmHg, respectively [46]. The channel activities are also affected by other factors, such as pH, temperature, membrane potential, arachidonic acid, and G protein-mediated regulation [43, 44, 47, 48, 49]. TRAAK and TREK give rise to leak (also called background) K+ currents. A well-known role of background K+ currents is to stabilize the negative resting membrane potential and counterbalance depolarization. TRAAK was recently reported to open upon ultrasound, and the single-channel conductance was ∼73pS [50]. Without the stimulation, TRAAK had an average open probability of 1.9%, while ultrasound and pressure increased channel open probability to 6.3 and 26%, respectively.

TRAAK has been reported in the mouse [47, 51] and salamander photoreceptors [30]. Mouse photoreceptors also express TREK1 [52]. Our team applied pressure on the rod IS and evoked sustained currents of three components in rods [30]. The pressure-induced outward current at membrane potentials ≥ −80 mV (termed Ipo) was sensitive to the blockage of intracellular Cs+, in line with the expression of MSCs permeable to K+. Using tetraethylammonium (TEA) to block most potassium channels largely depolarized salamander rods (33 mV, 82.5%) and reduced the light response by 24% [30].

Fink and colleagues cloned and expressed TRAAK in Xenopus oocytes and COS cells (fibroblast-like cell lines derived from monkey kidney tissue) and first reported the instantaneous and non-inactivating currents of TRAAK. The currents were not gated by voltage, only partially inhibited by Ba2+ at high concentrations, and were insensitive to the other classical K+ channel blockers TEA, 4-aminopyridine, and Cs+. TRAAK could be stimulated by the neuroprotective drug riluzole [47, 51], aprepitant [53], arachidonic acid (AA), and other unsaturated fatty acids but not by saturated fatty acids [47]. Wang and colleagues [51] explored the expression of TRAAK and the apoptosis of the ONL with immunostaining, Western blotting, and real-time polymerase chain reaction (RT-PCR). The channel agonist riluzole activated TRAAK and delayed the apoptosis of photoreceptors in the ONL of rd1 mice of retinal degeneration (Pde6brd1). They pointed out that activation of TRAAK in rd1 mice protects photoreceptors from apoptosis.

TRAAK, TREK1, and BK are permeable to K+, and a typical reversal potential (Erev) for K currents is around −75 to −90 mV [50, 54]. TREK is sensitive to the blockage of Cs+ [55], and TREK1 is inhibited by fluoxetine, norfluoxetine, and L-methionine [48, 56]. TREK2 is a 78% homology of TREK1. TREK2 produces rapidly activating and non-inactivating outward rectifier K+ currents. The single-channel conductance is 100 pS. The currents can be strongly stimulated by lysophosphatidylcholine, riluzole, polyunsaturated fatty acids such as arachidonic, docosahexaenoic, and linoleic acids, as well as volatile general anesthetics like chloroform, halothane, and isoflurane [41].

3.2 TRPs permeable to Na+ and Ca2+

TRPs are involved in various medical conditions (reviewed by Refs. [57, 58]) related to the bone [59], lung [60], kidney [61, 62, 63], digestive system [64, 65], pain [66, 67, 68, 69, 70], tooth, scapuloperoneal spinal muscular atrophy, hereditary motor and sensory neuropathy type IIC (HMSN IIC, also known as Charcot-Marie-Tooth disease type 2C (CMT2C)) [71, 72], Bardet-Biedl syndrome [73], hyponatremia [74], etc. TRPs include seven subfamilies, namely TRPC (canonical), TRPV, TRPM (melastatin), TRPN (NOMPC), TRPA (ANKTM1), TRPP (polycystin), and TRPML (mucolipin) [75, 76]. TRPs share the common feature of six transmembrane domains and various degrees of sequence similarity and permeability to cations.

The retina expresses multiple types of TRPs, but not all of them are known as mechano-gated channels. TRPs are variably modulated by membrane tension, osmolality, temperature, phorbol esters, and G-protein-mediated regulation. TRPA1, TRPC1/3/5/6; TRPM3/4/7, TRPP1/2, and TRPV2/4 are activated by stretch, osmolality, and/or pressure and identified as mechano-gated channels (reviewed by Refs. [77, 78, 79, 80]).

TRPVs include TRPV1-6. Their permeability to calcium (PCa) is higher than that to sodium (PNa). PCa/PNa is 9.6 (vanilloids) /3.8 (heat) for TRPV1, 3 for TRPV2, 2.8 for TRPV3, 6 for TRPV4, and > 100 for TRPV5 and 6 [77]. TRPV2 is activated by membrane tension, heat ≥52°C, cannabidiols, and 2-aminoethoxydiphenyl borate (2-APB) [81, 82, 83, 84, 85, 86], and piperlongumine is recently identified as a selective, reversible, and allosteric antagonist of human TRPV2 (hTRPV2) [87]. TRPV4 opens by mechanical pressure [88], membrane stretch [89], osmotic pressure [61, 90, 91, 92], warm temperature (27–34°C) [77, 93, 94, 95, 96], specific pharmacological agonists like GSK1016790A (GSK101, 1-2uM) and 4α-phorbol 12,13-didecanoate (4aPDD, 1-2uM), and other molecules [97], and it is inhibited by RN-1734 (∼20uM) [76, 98, 99, 100]. TRPV1 involves pressure-induced retinal pathology [101] and hypertonicity sensing in the primary osmosensory neurons in the mouse organum vasculosum lamina terminalis [102]; however, TRPV1 unlike TRPV2 is restricted to a subset of mechanically insensitive cutaneous nociceptors responding to heat [103]. TRPV1 is activated by noxious heat of 40–50°C [81, 104, 105], capsaicin, and resiniferatoxin [106, 107, 108], and it is selectively inhibited by capsazepine and 5′-iodoresiniferatoxin [107] and variably blocked by divalent heavy metals [109].

TRPs have been found in all retinal layers [110]. Photoreceptors, ONL, and OPL express TRPV1, TRPV2, TRPV4, TRPM7, TRPP2, TRPC1, and TRPML1 in the vertebrate and mammalian retina [30, 110, 111, 112]. TRPV1/VR1 immunoreactivity in goldfish and zebrafish retinas was present in photoreceptor synaptic ribbons [113]. TRPV2 was present in axons of photoreceptors and outer plexiform layer (OPL) in the rat, cat, primate [111], mouse [110], and salamander [30] retinas, and TRPV4 was observed in the OPL in the mouse [101, 114, 115] and primate retinas [54], and axons of salamander photoreceptors [30].

There are still disputes on whether the outer retinal neurons express TRPV4. In TRPV4 knockout mice, Yarishkin and colleagues [116] did not find significant changes in ERG a-wave and b-wave evoked by whole-field lights (presumably white) of 0.00025–79 cd.s/m2, and they concluded that TRPV4 did not regulate the distal retinal light responses. Several studies reported TRPV4 in retinal Müller cells, microglia cells, or astrocytes to critically involve volume regulation and swollen-related pathologies [117, 118, 119], indicating the importance of TRPV4 in glial cells. On the other hand, TRPV4 signals show a horizontal distribution pattern in the OPL in the mice, porcine, primate, and salamander retinas [30, 54, 110, 115], which are not well consistent with the vertical orientation of Müller cells and dispersed distribution of microglial cells. In mice of acute retinal detachment, the number of apoptotic photoreceptors was reduced by approximately 50% in TRPV4 knockout mice relative to wild-type mice [120].

TRPVs carry cation currents that reverse at ∼ −10 to 20 mV [30, 54, 85, 121], and the channel opening can cause membrane depolarization, regulate neuronal excitability, and mediate excitotoxicity [122, 123, 124]. Our recent data revealed the expression of TRPV4 and TRPV2 in the axon terminal of rods and cones and pressure-evoked currents and potentials in salamander photoreceptors [30]. We applied positive pressures onto the rod IS and elicited three components of currents in salamander rods, and one of them was a cation current that reversed at ∼ − 10 mV (termed Ipc). Meanwhile, hypotonicity induced a slow cation current of a similar reversal potential [30]. The data aligns with the immunocytochemical data and is consistent with the expression of TRPVs in photoreceptors. Also, pressure applied to the outer segment (OS) of rods and cones closed a Ca2+-dependent cation conductance reversed at ∼0 mV, in line with the closing of TRPV2 in the OS [30]. The wild-type rat TRPV2 was reported to be constitutively fully open [125], and the spatial structure of the agonist-free full-length TRPV2 molecular [126] showed larger upper and lower gates than the agonist-opened TRPV1. The pressure response in photoreceptors saturates at 25.9 mmHg. The agonists and antagonists of TRPV2 and TRPV4 variably affect the light response of salamander rods [30]. The light-evoked potentials recorded at various light intensities were larger at 23°C than at 31°C and severely disrupted at 43°C. The results indicate that mechanical stimuli may affect light signals in outer retinal neurons via TRPVs.

Besides, retinal pigment epithelium (RPE) is vital for the integrity of photoreceptors through its phagocytic function, and the RPE of the porcine, mouse, and human expresses TRPV2 [127, 128, 129]. The expression is enhanced by cannabidiol and inhibited by SKF96365 [128], a nonselective blocker of TRPs and inhibitor of the store-operated Ca2+ entry (SOCE). TRPV2 in RPE mediates both the heat-dependent and insulin-like growth factor-1-induced secretion of vascular endothelial growth factor (VEGF) [129]. RPE also expresses TRPV1-4, TRPMs, and TRPCs in the human retina [129, 130], TRPM3 in the mouse retina [110], and ENaCα in the rat and human retinas [131]. In freshly isolated RPE cells and ARPE-19 cells, Cordeiro and colleagues found that raising the temperature increased the secretion rate of VEGF-A. This effect was sensitive to ruthenium red (20uM). Heat also induced a Ca2+ signal, which was diminished by the TRPV blockers La3+ and ruthenium red and enhanced by 2-APB (3 mM) but insensitive to the TRPV1 agonist capsaicin and TRPV3 activator camphor.

3.3 ENaC and piezo permeable to Na+ or Na+/Ca2+

ENaC is a heteromultimeric channel usually composed of three homologous subunits (α, β, and γ) with a 30–40% identity at the level of their amino acid sequence [132, 133]. The immunoreactivity and mRNA of the pore-forming α-subunit have been observed in rods and cones of rat and human retinas [131, 134]. The specific blocker amiloride [135] enhanced ERG a-wave in the rat and rabbit retinas. Unlike other family members, ENaC is constitutively active [133, 136], which, in line with the observation in photoreceptors [135], suggests that ENaCα may regulate visual signals in photoreceptors in physiological conditions.

ENaC belongs to the ENaC/degenerin (ENaC/DEG) family and involves the functions of sensing and responding to mechanical and chemical stimuli [133, 137, 138]. It is regulated by the intracellular Na+ and H+, peptidases, and other small molecules [139, 140, 141, 142, 143, 144]. It is highly Na+-selective and amiloride-sensitive (EC50 150 nM). The permeability (P) ratio of PNa+: PK+ = 100: 1. ENaC/DEG is characterized by a relatively long extracellular loop bounded by two transmembrane pore-forming helices (TM1 and TM2). The single-channel conduction is 7.4pS in low-Na+ solution and 7.5pS in high-Na+ solution (reviewed by Refs. [133, 144]). AP 301 is an agonist of ENaCα [136, 145, 146, 147] that has been trialed for treating pulmonary permeability edema [148]. These results indicate that MSCs may open in physiological conditions to affect visual signals.

About one decade ago, Coste and colleagues [149, 150] revealed a novel family of mechanically activated cation channels in eukaryotes, consisting of Piezo1 and Piezo2 channels [149, 150, 151, 152]. Piezo senses touch and stretch in cells of fruit flies and mammals [79, 152, 153, 154, 155]. Piezo is the nonselective cationic channel that is permeable to alkali ions (K+, Na+, and Cs+), divalent cations (Ba2+, Ca2+, Mg2+, and Mn2+), and several organic cations (tetramethyl ammonium (TMA), TEA, etc.). The chord conductance of human Piezo 1 [156] at the membrane potential of -100 mV is 35-55pS for Cs+, Na+, and K+ and ∼ 23pS for Li+, and the conductance for divalent at 90 mM extracellular concentration is 25pS for Ba2+ and 15pS for Ca2+ at -80 mV and 10pS for Mg2+ at -50 mV measured at the cell-attached mode. Like TRPVs, Piezo carries cation currents, which reverse around −10 to 20 mV [18, 157]. The channel opening is anticipated to depolarize the membrane in physiological conditions and involve excitotoxicity in pathological conditions [158].

Piezo1 mRNA was detected in the mouse ONL [159]. Immunocytochemical and electrophysiological data from Bocchero and colleagues identified Piezo 1 and TRPC1 in rods of Xenopus retinas [160]. Mechanical stimulation in the order of 10pN applied to OS or IS evoked a “calcium transient,” and the channel blocker reduced the duration of the photo response to bright flashes. More interestingly, bright flashes of light caused a rapid shorting of OSs. The authors proposed that MSCs, including TRPC1 and Piezo, play an integral role in rod phototransduction in the vertebrate retina.

Advertisement

4. The expression and activity of MSCs in HCs

The OPL consists of the horizontally ramified axon terminals of photoreceptors, dendrites of BCs, and processes of HCs, and the IPL is composed of the axon terminals of BCs, processes of ACs, and dendrites of RGCs. The somas of BCs, ACs, and Müller cells reside in INL, BCs in the upper half of the INL, ACs in the lower half, and Müller cells in the middle [161]. Some earlier works revealed moderate loss of neurons in INL in glaucoma and pressure-caused retinal pathology in the outer retina. It has been unclear how MSCs are expressed in retinal interneuron BCs, HCs, and ACs.

In dissociated HCs from the retinas of the rat and mouse [162], Sun and colleagues first identified BK-mediated outward current in HCs. The single-channel conductance measured in symmetrical 150 mM K+ in mouse HCs was ∼250 pS (202–279 pS). BK-mediated membrane current was identified by the blockage of BK antagonists iberiotoxin (100 nM) and paxillin (2.5 mM), as well as Ca2+-free solutions, divalent cation, and voltage-gated calcium channel blockers. The potassium current was outwardly rectified and reversed around -75 mV. Blocking BK with paxillin depolarized the membrane and produced oscillation of increasing frequency, and the synthetic BK agonist NS1619 inhibited these oscillations. The authors concluded that the activation of BKa channels put a ceiling on membrane depolarization and regulated the temporal responsivity of HCs. The finding is consistent with the notion that BK hyperpolarizes HCs, counterbalancing the membrane depolarization and excitatory visual noises. In salamander retinas [30], the TRPV2 antibody brightly labeled the OPL, and half of the calretinin/GABA (gamma-aminobutyric acid)-positive processes of horizontal cells (HCs) were labeled for TRPV2. Pressure applied on the IS of salamander rods could evoke a Co2+-sensitive inward current component in rods at membrane potentials <−50 mV (termed Ipi). The data indicates the presence of TRPs in a presynaptic site, likely the processes of HCs.

Besides, D2 mice with congenital glaucoma showed thinning of OPL and loss of processes of BCs and HCs [163, 164].

Advertisement

5. The expression and activity of MSCs in BCs

BK was first observed in isolated ON BCs from the goldfish retina [27, 165] to mediate resonance (60–70 Hz or 5–10 Hz) K+ current sensitive to 100 nM charybdotoxin [165]. The potassium current was coupled with L-type voltage-gated calcium channels (L-CaV). BCs express TRAAK [47] and BK [166] in the mouse and rat retinas, TRPV4 in primates [54], and ENaC in rats [134]. Mammalian rod BCs (RBCs) express TRPV4 [54] and BKs [33, 166]. Nagai and colleagues used situ hybridization and immunohistochemistry and observed BK in rabbit RBCs and cone ON BCs [33]. In BKa mutant mice, ERG b-wave at the mesopic range (high scotopic range) is reduced, while others were not affected, indicating that BK involves the function of ON BCs [166, 167].

Dendrites of BCs ramify in the OPL, and OPL expresses BK [27, 29, 31], TRAAK [47], TREK1 [52], ENaC [168], TRPC1/4 [169], TRPV4 [114, 115], TRPV1 [113], and TRPV2 [30, 110, 111]. In the mouse retina, in situ hybridization and immunofluorescence revealed widespread expression throughout multiple retinal layers for TRPC1, TRPC3, TRPC4, TRPML1, and TRPP2 [110]. The somas of BCs reside in the INL, where TREK1, TREK2 [52], TRAAK [47, 52], ENaC [170], TRPV1 [106], and TRPV2 [111] have also been observed in mammals. The IPL was positive for TRAAK [30, 171], TRPV2 [111], TRPV4 [54, 114, 115], TRPC1/4 [169], and ENaCa [131, 168] in mammals and vertebrate retinas.

The dendrites and somas of BCs in the primate retina are positive for TRPV4 [54]. Pressure evoked transient cation currents in primate BCs, which reversed at ∼ −10 mV and were enhanced upon heating from 24° to 34° in line with TRPV4. The pressure for the half-maximal effect in the primate retina was ∼20 mmHg [54], comparable to the borderline of harmful IOP level in glaucoma. The data indicate that BCs may respond to the fluctuation of IOP levels in physiological and pathological conditions. Besides, TRPV1 was observed in the excitatory synapses from BCs in the dendrites of RGCs [172].

The mRNA and immunoreactivity of ENaCα have been observed in BCs in human and rat retinas [131, 134]. The function of BCs is accessible in vivo with ERG. In the presence of ENaC blocker amiloride, Brockway and colleagues [135] found that ERG a-, b- and d-waves were all enhanced in the rat retina, while the slow PIII-Müller cell response [135] was reduced. The data indicate that higher activities of ENaC can reduce the function of ON BCs in normal conditions. Besides, Piezo1 mRNA was detected in the mouse INL [159].

D2 mice of congenital glaucoma have shown thinning of OPL, loss of processes of BCs and HCs [163, 164], and reduction of ERG a- and b-waves. High IOP can result in the disturbance to the IPL [173] and reduction of BC synapses in RGC dendrites before RGC death [174, 175, 176] and RBC dendrites [173]. Pressurized airwaves can reduce ERG b-wave at post-blast 3–7 days [177, 178] and damage dendrites of rod BCs (RBCs) [179]. Interestingly, in studies of TRI, the data on ERG b-wave observed within days [177, 178] are inconsistent with those obtained in months [180, 181], resembling glaucoma [182, 183, 184]. ERG b-wave measures a field potential from DBC populations, which hardly measure activities of HBCs due to the very short duration of flash stimuli (∼5 ms for animals and 0.1–0.2 sec for patients [185, 186]). ERG a-wave [184, 187, 188] may reduce in glaucoma, and b-wave may reduce [163, 188, 189] or briefly increase in earlier stages of IOP elevation [184]. Besides potential experimental errors, the discrepancy in b-wave could indicate two progressive pathophysiological stages; the early one likely involves more closely to the responses of MSCs in BCs to initial IOP elevation, and the later one may be more attributable to damage to the structural integrity of BCs.

In a mouse model, the expression of PKCα was decreased progressively in cell bodies and dendrites of RBCs by the acute ocular hypertension (IOP ∼100 mmHg) induced by injection of physiological saline [190]. In the retinas injured by the intravitreal injection of 10 mM NMDA (N-methyl-D-aspartate), a patch-clamping study showed that RBCs were more vulnerable to excitotoxicity than cone BCs and that PKCα critically involved the function of mGluR6. The study pointed out that in the three mice models (ocular hypertension, excitotoxic neurodegeneration, or optic nerve crash), RBC dysfunction all occurs before RGC loss. Similarly, in a glaucoma model rat, the sodium hyaluronate-induced ocular hypertension caused damage to the OPL and IPL and was associated with morphologic and morphometric changes in BCs [173].

MSCs permeable to Na+ and Ca2+ respond to pressure with excitatory currents, and their role in pressure-induced glutamate release from photoreceptors and BCs requires more attention and further exploration.

Advertisement

6. The expression and activity of MSCs in ACs

BK is in the open state and mediates small outward miniature currents at −60 to −40 mV in salamander ACs, which were sensitive to iberiotoxin [191]. A17 ACs from rat retinas provide reciprocal inhibitory synapses to RBCs that mediate the primary rod pathway. BK reduces GABA release from A17 ACs and regulates the flow of excitatory signals through the primary rod pathway [192]. TRPC5 is present in mouse ACs and RGCs, serving as a negative regulator of RGCs for axon outgrowth [193].

Starburst ACs (SACs) are cholinergic GABAergic ACs. They critically modulate the activities of the direction-selective RGCs and mediate the retinal wave during development. Fort and colleagues [194] observed a strong expression of the KCNK2 gene that encodes TREK1 in isolated SACs from the mouse retina. At the postnatal day P1-6, TREK1−/− mice exhibited an altered frequency of the retinal wave, which is known to be set by the slow afterhyperpolarization (sAHP) of SACs. The sAHP conductance was found to be calcium-dependent, reversed at potassium reversal potential (EK), blocked by barium, insensitive to apamin and TEA, and activated by arachidonic acid.

Mammalian ACs express the TRAAK and TRPV2 transcripts [47, 111]. The INL shows the immunoreactivity of TRAAK, TREK1, and TREK2 [52]. A low level of TRAAK is present in the IPL of the salamander retina [30]. The mRNA and immunoreactivity of ENaCα were increased in the IPL and INL in D2 mice [168], and the INL of human retinas expressed the mRNA of β- and γ-subunits of ENaC [170]. These studies suggest that individual interneurons require several types of MSCs to perform normal functions.

It has been poorly understood how ACs express MSCs and respond to the mechanical activation of MSCs. ACs mediate spontaneous postsynaptic currents (sIPSCs) in RGCs and BCs. In rat RGCs, TRPV4 antagonist HC-067047 reduces sIPSCs, and TRPV4 agonist GSK 1016790A enhances sIPSCs in ON and OFF RGCs. TRPV4 activation has different effects on the release of glycine and GABA and the function of the neurotransmitter receptors [195]. TRPV1 was found in processes and perikarya of putative ACs in the INL, which was in the vicinity of nNOS-positive ACs but did not colocalize nNOS, and retinal explants exposed to capsaicin presented high protein nitration and cell death in the INL and GCL [106].

ACs have shown some changes in glaucoma retinas. Pressure insult led to Type III neuron death restricted to the INL [196] in an early report. Later studies in glaucoma models further identified significantly [25, 197] or minor loss (∼15%) of ACs [198, 199], low light sensitivity of AII ACs [184, 200], thinning of OPL, loss of processes of BCs and HCs [163, 164], disturbance to the IPL [173], and reduction of BC synapses in RGC dendrites [174, 175, 176, 201] before RGC death. AII ACs are glycinergic ACs, making ∼11% of all ACs [202] in the rabbit retina. GABAergic ACs, including cholinergic ACs [203], were also reduced in glaucoma model D2 mice [198] and rats with elevated IOP [204]. It was concurrent with the RGC death and appeared to be selective for ACs coupled to RGCs [204]. Different changes were observed in nitroxidergic (NO) ACs [198] in glaucoma models. IOP elevation in rats reduced the glutamate- and K+-induced GABA release and increased GABA uptake [205]. In a glaucoma model rat, the sodium hyaluronate-induced ocular hypertension [173] damaged the direction-selective circuit, IPL, and OPL besides RGC populations. The treated eyes exhibited morphologic and morphometric changes in BCs, ON-OFF direction-selective RGCs, ON and OFF SACs, and the IPL. The data demonstrated the variation of pressure responsiveness among ACs.

Advertisement

7. The expression and activity of MSCs in RGCs

7.1 MSCs permeable to K+ in RGCs

BK has been observed in RGCs from mammalian and other vertebrate retinas [191, 206, 207, 208]. It is in open states and involves the development and normal function of RGCs in vertebrate retinas. In isolated trout retinal RGCs [208], Iberiotoxin-sensitive, calcium-activated potassium currents mediated by BK were minimal before hatching but increased significantly then.

In the isolated and intact RGCs [207] from ferret retinas, charybdotoxin, a blocker of BK, increased RGC spiking. Charybdotoxin and apamin (SKCa channel blocker) reduced the time to the threshold and the hyperpolarization after the spike in isolated RGCs and 80% of a- and b-RGCs. In mouse RGCs, blocking BK with charybdotoxin increased the spontaneous excitatory post-synaptic currents (EPSCs) and light-evoked ON-EPSCs but decreased the light-evoked OFF-IPSCs [206]. These results align with the notion that BK regulates the membrane potential, excitability, and visual signals of mammalian RGCs.

The GCL of the mouse retina exhibits immunoreactivities of TREK-1, TREK2 [52, 171], and TRAAK [47, 171]. Hughes and colleagues also observed other K2P channels in RGCs, such as TASK-1, TWIK-1, TWIK-2, and TWIK-3 [52]. In rat RGCs, TREK2 at the sites postsynaptic to GABAergic ACs may be activated by GABAB receptors to affect the activities of RGCs [209]. Using immunocytochemistry, patch-clamp, PCR, and western blot, Zhang and colleagues [210] found that rd1 mice express higher levels of the mRNA and protein of TREK1 and TRAAK, and the arachidonic acid-evoked current was bigger in RGCs. TASK-3 also showed some regulation on the excitability of mouse RGCs [211]. The data suggest that K2Ps have physiological and pathological roles in RGCs.

7.2 Piezo and ENaC in RGCs

Piezo proteins are pore-forming mechano-gated channels [150, 154]. Optic never head expresses multiple MSCs [212]. The mRNA of Piezo1 and Piezo2 are present in the myelinated region of the mouse optic nerve, and the expression level of Piezo2 is high in the optic never head. Immunostaining revealed Piezo1 and Piezo2 in the GCL [213] and non-neuronal ocular tissues [213, 214], as well as astrocytes [215]. Zhu and colleagues used single-molecule fluorescence in situ hybridization (smFISH) and transgenic reporter mice expressing Piezo fusion proteins [159] to explore the distribution of Piezo in the mouse retina. Piezo1 and Piezo2 were found in the GCL, trabecular meshwork, and ciliary body, and Piezo1 mRNA was more abundant. In genetically encoded Ca2+ indicator mice and an ex vivo pressurized retina preparation, Harraz and colleagues reported Ca2+-permeable Piezo1 in the endothelial cells in retinal and cortical capillaries [214].

Piezo channel plays a part in RGC damage in the mouse retina. The expression of retinal Piezo2 increases in the mouse model of high IOP. Piezo1 agonist Yoda 1 suppressed neurite outgrowth in RGCs. On the other hand, Piezo antagonist GsMTx4 promoted neurite outgrowth in RGCs [213, 216]. A gain of function variant of Piezo1 (e756del) was observed in 30% of African Americans and associated with higher IOPs, thinner NFL, and lower optic nerve head capillary densities, but these did not reach statistical significance [158]. Piezo1 regulates the cell cycle progression of astrocytes, and loss of Piezo1 can lead to the cell cycle of astrocytes at the optic nerve head arrest at the G0/G1 phase [215].

ENaC is a novel therapeutic target for serval human diseases [133]. It is activated by protease and blocked by amiloride [136, 144]. ENaCa immunoreactivity and mRNA expression are present in the GCL and IPL in the mouse retina [168] and RGCs in the rat and human retina [131, 134]. GCL expresses the mRNA and protein of β- and γ-subunits of ENaC in the human retina [170]. In D2 mice, Dyka and colleagues found an upregulation of ENaCα gene expression in the IPL and GCL, but they did not find β- and γ-subunits [168]. A recent study reveals that mobile Zn2+ from interneurons triggers RGC death in optic nerve injury [217], while Zn2+ may inhibit TRPM1 and activate TREK2 [218, 219].

7.3 TRPVs in RGCs

The retina expresses multiple types of TRPs of variable functional significance (reviewed by Refs. [57, 58]). Gilliam and colleagues identified TRPs in the mouse retina with RT-PCR and immunohistology [110]. The strongest signals were reported for TRPC1, TRPC3, TRPM1, TRPM3, and TRPML1, and clear signals were obtained for TRPC4, TRPM7, TRPP2, TRPV2, and TRPV4. Their results revealed widespread expression throughout multiple retinal layers for TRPC1, TRPC3, TRPC4, and TRPML1. Other early studies have described the distribution patterns of multiple TRPs in a variety of species [1, 101, 114, 115, 220]. RGCs and their axons have been widely identified as the most vulnerable to pressure in both TRI [5, 221, 222] and glaucoma [3, 4], and recent data from glaucoma models have revealed RGC dysfunction [200] and synaptic damage [174, 223, 224] before axon loss. Yet, the involvement of MSCs has been uncertain.

GCL and IPL express TRPV2 in rat, cat, primate, and salamander retinas [30, 54, 106, 111, 112], TRPV1 in rat and primate retina [101, 106], and TRPV4 in the mouse [101, 114], porcine [115], and primate retina [54]. The optic nerve head was revealed mRNAs of TRPV2 and TRPV4 [212]. Lakk and colleagues observed mRNAs of TRPVs in isolated RGCs of 7–15 um diameter from the mouse retina, whose levels were TRPV4 > TRPV2 > TRPV3 and TRPV1 [225].

TRPV4 single-channel conductance was measured as an inward conductance of 60pS and outward conductance of 102pS. TRPV3 showed a higher single-channel conductance, 172 pS at +60 mV, compared with ∼100pS for TRPV1/2/4 [121]. Liedtke and colleagues first cloned cDNAs encoding the vanilloid receptor-related osmotically activated channel (VR-OAC, i.e., TRPV4) from the rat, mouse, human, and chicken. They showed that TRPV4 was a cation channel gated by exposure to hypotonicity within the physiological range [89]. Later, the researchers confirmed the pressure responsiveness of TRPV4 in transfected cell lines and TRPV4 mutant mice [61, 92]. Another group of scientists [88] separately reported the impaired osmotic sensation in mice lacking TRPV4.

The large peripheral RGCs of the primate retina express more heavily TRPV4 [54]. RGCs in the mouse [114] and primate retina [54] can be activated by micromolar TRPV4 agonists GSK1016790A and 4aPDD, exhibiting membrane depolarization and higher firing rate. In cultured RGCs, TRPV4 agonists evoked calcium influxes and were associated with apoptosis of the neurons [114]. TRPV4 antagonist RN-1734 has been tested in retinal slices in culture and revealed a neuroprotective role in the porcine retina [115]. These observations have confirmed the expression of TRPV4 in RGCs. TRPV4 agonist enhances the frequency of excitatory postsynaptic currents consistent with the expression of TRPV4 in presynaptic BCs [54]. Notably, activation of TRPV4 can increase the excitability of RGCs via shortening the delay [54] and augmenting the amplitude [226] of voltage-gated sodium current sensitive to tetrodotoxin. Li and colleagues [118] reported an increased expression of TRPV4 under high IOP. Intravitreal injection of TRPV4 agonist induces Müller cell gliosis, and activation of TRPV4 induces the release of tumor necrosis factor-α (TNF-α) from cultured Müller cells. Inhibition of TNF-a could reduce TRPV4-mediated RGC apoptosis [118]. The results together support that the normal functions and apoptosis of RGCs involve TRPV4 located in RGCs, BCs, and Müller cells. In Müller cells, TRPV4 serves as a volume regulator [117, 120, 226, 227, 228, 229].

TRPVs also involve diabetic retinopathy. TRPV4 knockout or inhibition could prevent the increased water diffusion and blood-retina barrier breakdown in the retina of streptozotocin-induced diabetic mice [230]. In rat models, TRPV2 contributes to stretch-activated cation currents and myogenic constriction in retinal arterioles, and genetic deletion of TRPV2 impaired the myogenic reaction of retinal arterioles, resembling that observed in diabetic animals [85, 231]. Systemic activation of TRPV4 could cause endothelial failure and circulatory collapse [60], and TRPV4 promotes vascular permeability in retinal vascular disease [232]. Calcium influx through TRPV4 modulates the adherens contacts between retinal microvascular endothelial cells [233], and pharmacologic inhibition of TRPV1 or TRPV4 suppresses retinal angiogenesis in vitro [234].

RGCs express TRPV1 [101, 108, 172, 235, 236, 237, 238]. Sappington and colleagues reported mRNA of TRPV1 in the cell body and axon of RGCs, and the TRPV1 level increased by IOP elevation in D2 mice. They applied hydrostatic pressure (70 mmHg) to RGCs in culture and reported that TRPV1 antagonism could reduce the pressure-induced RGC apoptosis. Similar results were obtained from RGCs of the monkey, human, and rat retina, and TRPV1 knockout and pharmacological antagonism of TRPV1 were found to prevent the pressure-induced RGC apoptosis [101]. TRPV1 and TRPV4 were shown to form a protein complex in RGCs [101], but the interaction between TRPV1 and TRPV4 appears to be insignificant [225]. In isolated mouse RGCs from wild-type and TRPV1−/− mice, Lakk and colleagues examined calcium influxes via TRPV4 by calcium imaging, and TRPV1 mutation showed no effect on TRPV4 activities evoked by the channel agonist GSK1016790 [225]. However, TRPV1 mutation and antagonist were reported to be neuroprotective for isolated RGCs [101] and retinal explants [101] but neurodegenerative in vivo [101, 108, 238], indicating the differential roles of TRPV1 in RGCs and presynaptic neurons.

In transgenic mice where a few subtypes of RGCs were genetically labeled, the RGCs ramified in the sublamina a showed the greatest change in the dendritic morphology after 1 week of IOP elevation in the microbead occlusion model of glaucoma [176]. On the other hand, both ONaGCs and OFFaGC exhibited reduced light sensitivity after a few weeks of IOP elevation, as shown in our results [200]. Rountree and colleagues applied pulsatile injection (pulse-width > 50 ms at 0.69 kPa pressure) of Ames mediums into retinal tissue and evoked neuronal responses comparable to light responses. The response of RGCs was reduced by a TRPV blocker [239]. Genetic deletion of TRPV1 can differently affect the excitability of RGCs firing continuously to light onset (ON-sustained) vs. light offset (OFF-sustained) [236]. These results together indicate that TRPVs may play physiological and excitotoxic roles in RGCs.

Advertisement

8. Other mechanosensitive TRPs in retinal neurons

TRPA1, TRPC1/3/5/6, TRPM3/4/7, and TRPP1 are mechano-gated channels (reviewed by Refs. [79, 80]). TRPA1, TRPM1, and TRPCs mediate light response in sensory neurons. Planarians possess extraocular photoreception and display an unconventional TRPA1-mediated photophobic response to near-UV light [240]. Glutamate released from photoreceptors hyperpolarizes ON BCs by stimulating mGluR6, which was mediated by the inactivation of TRPM1 after binding with the G-protein a or bg subunits [241, 242] (reviewed in Refs. [57, 243]) to mediate light signals. TRPC6 and TRPC7 are expressed in the intrinsic photosensitive RGCs (ipRGCs) in the mammalian retina to involve melanopsin phototransduction. The melanopsin photocurrent in ipRGCs was abolished by blocking or eliminating TRPC3/6/7 channels in M1-type ipRGCs and TRPC3/6 in M2-type ipRGCs, suggesting that the target of melanopsin phototransduction varies in these ipRGCs (reviewed in Ref. [244]). TRPC1/4/5/6 are expressed in endothelial and Müller cells, and the TRPC1/4/5/6(−/−) compound knockout mice showed resistance to diabetic retinopathy [245].

TRPC5 is present in mouse RGCs and ACs to serve as a negative regulator of RGCs for axon outgrowth [193]. Optic never head expresses multiple types of MSCs. The mRNA of TRPP1 and TRPP2 were observed in the myelinated region of the mouse optic nerve [212]. The optic nerve head and the glial lamina were revealed TRPC1-6, TRPV2, TRPV4, TRPM1-4/6/7, TRPP1, and TRPP2. About 43–87% of individual astrocytes express the mRNA of TRPC1, TRPM7, TRPP1, and TRPP2. The expression level of TRPP2 was high in the optic nerve head. TRPM3 immunofluorescence was present in a subset of RGCs in mice during postnatal 7–14 days and in adult mice. Activation of TRPM3 with the synthetic TRPM3 agonist CIM0216 (CIM) induced prolonged calcium transients in RGCs, which were mostly absent in TRPM3 mutant mice. The prolonged calcium transient was not associated with strong membrane depolarizations but induced c-Fos expression [246, 247].

Ischemia elicited a decrease in the ERG-recorded retinal responsiveness to light along with reactive gliosis and a significant increase in the expression of TRPM7 in Müller cells [248]. In another report [249], blue light triggered apoptosis of retinal pigment epithelial (RPE) cells, and its deleterious effects were partially attenuated by the synergistic action of TRPM7 and the pigment epithelium-derived factor (PEDF) via the PKC/ERK signaling pathway. The reverse transcription PCR analysis demonstrated the mRNA expression of TRPC1, TRPM7, TRPV1/2/4, and TRPP1, but not TRPC6 or TRPM4 in retinal vascular smooth muscle cells (VSMCs). TRPV2 inhibitor tranilast and specific TRPV2 pore-blocking antibodies reserved the hypoosmotic stretch-induced Ca2+ influx in VSMCs of isolated retinal arterioles, but inhibitors of TRPC1, TRPM7, TRPV1, and TRPV4 had no effect. The authors concluded that retinal VSMCs expressed a range of mechanosensitive TRP channels, but only TRPV2 contributed to myogenic signaling in this vascular bed [85].

Advertisement

9. Conclusions

Fluctuating mechanical stresses and strains are present in the ocular environment and are associated with multiple retinal disorders. Individual retinal neurons often co-express several types of MSCs, which may mediate either depolarizing or hyperpolarizing membrane currents upon mechanical and chemical activation. Each MSC contributes more-or-less to the membrane currents or potentials, regulating neuronal excitability and visual signals or resulting in neuronal excitotoxicity. A few TRPs directly mediate phototransduction in ipRGCs and synaptic transmission in ON BCs in mammalian retinas, and some of them are also mechano-gated channels. MSCs are still to be fully characterized for their significance in maintaining cellular mechanical homeostasis, aging change, and cooperation in physiological and pathological conditions, facilitating the development of effective therapeutic strategies for pressure-related visual disorders.

Advertisement

Acknowledgments

I thank the funding support from US Army Medical Research Acquisition Activities VR210010 and the NIH core grant EY02520.

Conflict of interest

The author declares no conflict of interest.

References

  1. 1. Pang JJ. Roles of the ocular pressure, pressure-sensitive ion channel, and elasticity in pressure-induced retinal diseases. Neural Regeneration Research. 2021;16:68-72
  2. 2. Conlon R, Saheb H, Ahmed II. Glaucoma treatment trends: A review. Canadian Journal of Ophthalmology. 2017;52:114-124
  3. 3. Jonas JB, Aung T, Bourne RR, Bron AM, Ritch R, Panda-Jonas S. Glaucoma. Lancet. 2017;390:2183-2193
  4. 4. Quigley HA. Glaucoma. Lancet. 2011;377:1367-1377
  5. 5. Sen N. An insight into the vision impairment following traumatic brain injury. Neurochemistry International. 2017;111:103-107. DOI: 10.1016/j.neuint.2017.01.019. Epub 2017 Feb 2
  6. 6. Mufti O, Mathew S, Harris A, Siesky B, Burgett KM, Verticchio Vercellin AC. Ocular changes in traumatic brain injury: A review. European Journal of Ophthalmology. 2020;30:867-873
  7. 7. Thomas CN, Courtie E, Bernardo-Colon A, Essex G, Rex TS, Ahmed Z, et al. Assessment of necroptosis in the retina in a repeated primary ocular blast injury mouse model. Experimental Eye Research. 2020;197:108102. DOI: 10.1016/j.exer.2020.108102. Epub 2020 Jun 6
  8. 8. Curcio CA, Sloan KR, Kalina RE, Hendrickson AE. Human photoreceptor topography. The Journal of Comparative Neurology. 1990;292:497-523
  9. 9. Jonas JB, Schneider U, Naumann GO. Count and density of human retinal photoreceptors. Graefe's Archive for Clinical and Experimental Ophthalmology. 1992;230:505-510
  10. 10. Rodieck RW. The First Steps in Seeing. Sunderland MA: Sinauer Associates; 1998
  11. 11. Ortin-Martinez A, Nadal-Nicolas FM, Jimenez-Lopez M, Alburquerque-Bejar JJ, Nieto-Lopez L, Garcia-Ayuso D, et al. Number and distribution of mouse retinal cone photoreceptors: Differences between an albino (Swiss) and a pigmented (C57/BL6) strain. PLoS One. 2014;9:e102392
  12. 12. Jeon CJ, Strettoi E, Masland RH. The major cell populations of the mouse retina. The Journal of Neuroscience. 1998;18:8936-8946
  13. 13. Rohlich P, Van Veen T, Szel A. Two different visual pigments in one retinal cone cell. Neuron. 1994;13:1159-1166
  14. 14. Jacobs GH, Neitz J, Deegan JF. Retinal receptors in rodents maximally sensitive to ultraviolet light. Nature. 1991;353:655-656
  15. 15. Applebury ML, Antoch MP, Baxter LC, Chun LL, Falk JD, Farhangfar F, et al. The murine cone photoreceptor: A single cone type expresses both S and M opsins with retinal spatial patterning. Neuron. 2000;27:513-523
  16. 16. Volland S, Esteve-Rudd J, Hoo J, Yee C, Williams DS. A comparison of some organizational characteristics of the mouse central retina and the human macula. PLoS One. 2015;10:e0125631
  17. 17. Liu C, Montell C. Forcing open TRP channels: Mechanical gating as a unifying activation mechanism. Biochemical and Biophysical Research Communications. 2015;460:22-25
  18. 18. Fang XZ, Zhou T, Xu JQ, Wang YX, Sun MM, He YJ, et al. Structure, kinetic properties and biological function of mechanosensitive piezo channels. Cell & Bioscience. 2021;11:13-00522
  19. 19. Martinac B. Mechanosensitive ion channels: Molecules of mechanotransduction. Journal of Cell Science. 2004;117:2449-2460
  20. 20. Khurana AK. Ophthalmology. New Delhi: New Age International; 2005
  21. 21. Goel M, Picciani RG, Lee RK, Bhattacharya SK. Aqueous humor dynamics: A review. The Open Ophthalmology Journal. 2010;4:52-59. DOI: 10.2174/1874364101004010052
  22. 22. Pietrowska K, Dmuchowska DA, Krasnicki P, Mariak Z, Kretowski A, Ciborowski M. Analysis of pharmaceuticals and small molecules in aqueous humor. Journal of Pharmaceutical and Biomedical Analysis. 2018;159:23-36. DOI: 10.1016/j.jpba.2018.06.049. Epub 2018 Jun 25
  23. 23. Kaufman P, Alm A, Levin LA, Nilsson SFE, Hoeve JV, Wu S. Adler's Physiology of the Eye. Philadelphia: Saunders/Elsevier; 2011
  24. 24. Herndon LW, Graul TS, Jones LS, Henderer JD, Khaimi MA, Mathers KJ, et al. Academy MOC Essentials Practicing Ophthalmologists Curricumum 2017-2019 Glaucoma. San Francisco, CA: American Academy of Ophthalmololgy; 2019
  25. 25. Pang JJ, Wu SM. Ocular pressure-volume relationship and ganglion cell death in glaucoma. OBM Neurobiology. 2021;5:10
  26. 26. Shah KR, Guan X, Yan J. Structural and functional coupling of calcium-activated BK channels and calcium-permeable channels within nanodomain signaling complexes. Frontiers in Physiology. 2022;12:796540. DOI: 10.3389/fphys.2021.796540. eCollection 2021
  27. 27. Sakaba T, Ishikane H, Tachibana M. Ca2+ −activated K+ current at presynaptic terminals of goldfish retinal bipolar cells. Neuroscience Research. 1997;27:219-228
  28. 28. MacLeish PR, Nurse CA. Ion channel compartments in photoreceptors: Evidence from salamander rods with intact and ablated terminals. Journal of Neurophysiology. 2007;98:86-95
  29. 29. Xu JW, Slaughter MM. Large-conductance calcium-activated potassium channels facilitate transmitter release in salamander rod synapse. The Journal of Neuroscience. 2005;25:7660-7668
  30. 30. Pang JJ, Gao F, Wu SM. Generators of pressure-evoked currents in vertebrate outer retinal neurons. Cells. 2021;10:1288
  31. 31. Pelucchi B, Grimaldi A, Moriondo A. Vertebrate rod photoreceptors express both BK and IK calcium-activated potassium channels, but only BK channels are involved in receptor potential regulation. Journal of Neuroscience Research. 2008;86:194-201
  32. 32. Mani BK, Brueggemann LI, Cribbs LL, Byron KL. Activation of vascular KCNQ (Kv7) potassium channels reverses spasmogen-induced constrictor responses in rat basilar artery. British Journal of Pharmacology. 2011;164:237-249
  33. 33. Nagai N, Koyanagi E, Izumida Y, Liu J, Katsuyama A, Kaji H, et al. Long-term protection of genetically ablated rabbit retinal degeneration by sustained Transscleral Unoprostone delivery. Investigative Ophthalmology & Visual Science. 2016;57:6527-6538
  34. 34. Qi Z, Chi S, Su X, Naruse K, Sokabe M. Activation of a mechanosensitive BK channel by membrane stress created with amphipaths. Molecular Membrane Biology. 2005;22:519-527
  35. 35. Jiang Y, Lee A, Chen J, Cadene M, Chait BT, MacKinnon R. Crystal structure and mechanism of a calcium-gated potassium channel. Nature. 2002;417:515-522
  36. 36. Zeng XH, Xia XM, Lingle CJ. Divalent cation sensitivity of BK channel activation supports the existence of three distinct binding sites. The Journal of General Physiology. 2005;125:273-286
  37. 37. Maingret F, Patel AJ, Lesage F, Lazdunski M, Honore E. Lysophospholipids open the two-pore domain mechano-gated K(+) channels TREK-1 and TRAAK. The Journal of Biological Chemistry. 2000;275:10128-10133
  38. 38. Patel AJ, Lazdunski M, Honore E. Lipid and mechano-gated 2P domain K(+) channels. Current Opinion in Cell Biology. 2001;13:422-428
  39. 39. Kim Y, Bang H, Gnatenco C, Kim D. Synergistic interaction and the role of C-terminus in the activation of TRAAK K+ channels by pressure, free fatty acids and alkali. Pflügers Archiv. 2001;442:64-72
  40. 40. Patel AJ, Honore E, Maingret F, Lesage F, Fink M, Duprat F, et al. A mammalian two pore domain mechano-gated S-like K+ channel. The EMBO Journal. 1998;17:4283-4290
  41. 41. Lesage F, Terrenoire C, Romey G, Lazdunski M. Human TREK2, a 2P domain mechano-sensitive K+ channel with multiple regulations by polyunsaturated fatty acids, lysophospholipids, and Gs, Gi, and Gq protein-coupled receptors. The Journal of Biological Chemistry. 2000;275:28398-28405
  42. 42. Cox CD, Bavi N, Martinac B. Origin of the force: The force-from-lipids principle applied to piezo channels. Current Topics in Membranes. 2017;79:59-96
  43. 43. Enyedi P, Czirjak G. Molecular background of leak K+ currents: Two-pore domain potassium channels. Physiological Reviews. 2010;90:559-605
  44. 44. Brohawn SG. How ion channels sense mechanical force: Insights from mechanosensitive K2P channels TRAAK, TREK1, and TREK2. Annals. New York Academy of Sciences. 2015;1352:20-32
  45. 45. Dedman A, Sharif-Naeini R, Folgering JH, Duprat F, Patel A, Honore E. The mechano-gated K(2P) channel TREK-1. European Biophysics Journal. 2009;38:293-303
  46. 46. Maingret F, Fosset M, Lesage F, Lazdunski M, Honore E. TRAAK is a mammalian neuronal mechano-gated K+ channel. The Journal of Biological Chemistry. 1999;274:1381-1387
  47. 47. Fink M, Lesage F, Duprat F, Heurteaux C, Reyes R, Fosset M, et al. A neuronal two P domain K+ channel stimulated by arachidonic acid and polyunsaturated fatty acids. The EMBO Journal. 1998;17:3297-3308
  48. 48. Kennard LE, Chumbley JR, Ranatunga KM, Armstrong SJ, Veale EL, Mathie A. Inhibition of the human two-pore domain potassium channel, TREK-1, by fluoxetine and its metabolite norfluoxetine. British Journal of Pharmacology. 2005;144:821-829
  49. 49. Brohawn SG, Del Marmol J, MacKinnon R. Crystal structure of the human K2P TRAAK, a lipid- and mechano-sensitive K+ ion channel. Science. 2012;335:436-441
  50. 50. Sorum B, Rietmeijer RA, Gopakumar K, Adesnik H, Brohawn SG. Ultrasound activates mechanosensitive TRAAK K(+) channels through the lipid membrane. Proceedings of the National Academy of Sciences of the United States of America. 2021;118:e2006980118
  51. 51. Wang L, Shi KP, Li H, Huang H, Wu WB, Cai CS, et al. Activation of the TRAAK two-pore domain potassium channels in rd1 mice protects photoreceptor cells from apoptosis. International Journal of Ophthalmology. 2019;12:1243-1249
  52. 52. Hughes S, Foster RG, Peirson SN, Hankins MW. Expression and localisation of two-pore domain (K2P) background leak potassium ion channels in the mouse retina. Scientific Reports. 2017;7:46085
  53. 53. McCoull D, Veale EL, Walsh Y, Byrom L, Avkiran T, Large JM, et al. Aprepitant is a novel, selective activator of the K2P channel TRAAK. Biochemical and Biophysical Research Communications. 2022;588:41-46. DOI: 10.1016/j.bbrc.2021.12.031. Epub 2021 Dec 14
  54. 54. Gao F, Yang Z, Jacoby RA, Wu SM, Pang JJ. The expression and function of TRPV4 channels in primate retinal ganglion cells and bipolar cells. Cell Death & Disease. 2019;10:364-1576
  55. 55. Huang D, Yu B. Recent advance and possible future in TREK-2: A two-pore potassium channel may involved in the process of NPP, brain ischemia and memory impairment. Medical Hypotheses. 2008;70:618-624
  56. 56. Lei Q, Pan XQ, Chang S, Malkowicz SB, Guzzo TJ, Malykhina AP. Response of the human detrusor to stretch is regulated by TREK-1, a two-pore-domain (K2P) mechano-gated potassium channel. The Journal of Physiology. 2014;592:3013-3030
  57. 57. Krizaj D, Cordeiro S, Straub O. Retinal TRP channels: Cell-type-specific regulators of retinal homeostasis and multimodal integration. Progress in Retinal and Eye Research. 2023;92:101114. DOI: 10.1016/j.preteyeres.2022.101114. Epub 2022 Sep 24
  58. 58. Yang TJ, Yu Y, Yang JY, Li JJ, Zhu JY, Vieira JAC, et al. Involvement of transient receptor potential channels in ocular diseases: A narrative review. Annals of Translational Medicine. 2022;10:839-6145
  59. 59. Krakow D, Vriens J, Camacho N, Luong P, Deixler H, Funari TL, et al. Mutations in the gene encoding the calcium-permeable ion channel TRPV4 produce spondylometaphyseal dysplasia, Kozlowski type and metatropic dysplasia. American Journal of Human Genetics. 2009;84:307-315
  60. 60. Willette RN, Bao W, Nerurkar S, Yue TL, Doe CP, Stankus G, et al. Systemic activation of the transient receptor potential vanilloid subtype 4 channel causes endothelial failure and circulatory collapse: Part 2. The Journal of Pharmacology and Experimental Therapeutics. 2008;326:443-452
  61. 61. Liedtke W, Friedman JM. Abnormal osmotic regulation in trpv 4−/− mice. Proceedings of the National Academy of Sciences of the United States of America. 2003;100:13698-13703
  62. 62. Cohen DM. The transient receptor potential vanilloid-responsive 1 and 4 cation channels: Role in neuronal osmosensing and renal physiology. Current Opinion in Nephrology and Hypertension. 2007;16:451-458
  63. 63. Wu L, Gao X, Brown RC, Heller S, O'Neil RG. Dual role of the TRPV4 channel as a sensor of flow and osmolality in renal epithelial cells. American Journal of Physiology. Renal Physiology. 2007;293:F1699-F1713
  64. 64. Holzer P. Transient receptor potential (TRP) channels as drug targets for diseases of the digestive system. Pharmacology & Therapeutics. 2011;131:142-170
  65. 65. Holzer P. TRP channels in the digestive system. Current Pharmaceutical Biotechnology. 2011;12:24-34
  66. 66. Alessandri-Haber N, Dina OA, Yeh JJ, Parada CA, Reichling DB, Levine JD. Transient receptor potential vanilloid 4 is essential in chemotherapy-induced neuropathic pain in the rat. The Journal of Neuroscience. 2004;24:4444-4452
  67. 67. Alessandri-Haber N, Dina OA, Joseph EK, Reichling DB, Levine JD. Interaction of transient receptor potential vanilloid 4, integrin, and SRC tyrosine kinase in mechanical hyperalgesia. The Journal of Neuroscience. 2008;28:1046-1057
  68. 68. Cao DS, Yu SQ, Premkumar LS. Modulation of transient receptor potential Vanilloid 4-mediated membrane currents and synaptic transmission by protein kinase C. Molecular Pain. 2009;5:5. DOI: 10.1186/1744-8069-5-5
  69. 69. Cortright DN, Szallasi A. TRP channels and pain. Current Pharmaceutical Design. 2009;15:1736-1749
  70. 70. Eid SR, Cortright DN. Transient receptor potential channels on sensory nerves. Handbook of Experimental Pharmacology. 2009;194:261-281
  71. 71. Deng HX, Klein CJ, Yan J, Shi Y, Wu Y, Fecto F, et al. Scapuloperoneal spinal muscular atrophy and CMT2C are allelic disorders caused by alterations in TRPV4. Nature Genetics. 2010;42:165-169
  72. 72. Landoure G, Zdebik AA, Martinez TL, Burnett BG, Stanescu HC, Inada H, et al. Mutations in TRPV4 cause Charcot-Marie-tooth disease type 2C. Nature Genetics. 2010;42:170-174
  73. 73. Tan PL, Barr T, Inglis PN, Mitsuma N, Huang SM, Garcia-Gonzalez MA, et al. Loss of Bardet Biedl syndrome proteins causes defects in peripheral sensory innervation and function. Proceedings of the National Academy of Sciences of the United States of America. 2007;104:17524-17529
  74. 74. Tian W, Fu Y, Garcia-Elias A, Fernandez-Fernandez JM, Vicente R, Kramer PL, et al. A loss-of-function nonsynonymous polymorphism in the osmoregulatory TRPV4 gene is associated with human hyponatremia. Proceedings of the National Academy of Sciences of the United States of America. 2009;106:14034-14039
  75. 75. Montell C. The TRP superfamily of cation channels. Science’s STKE. 2005;2005:re3
  76. 76. Nilius B, Szallasi A. Transient receptor potential channels as drug targets: From the science of basic research to the art of medicine. Pharmacological Reviews. 2014;66:676-814
  77. 77. O’Neil RG, Heller S. The mechanosensitive nature of TRPV channels. Pflügers Archiv. 2005;451:193-203
  78. 78. Clapham DE. SnapShot: Mammalian TRP channels. Cell. 2007;129:220
  79. 79. Barbeau S, Gilbert G, Cardouat G, Baudrimont I, Freund-Michel V, Guibert C, et al. Mechanosensitivity in pulmonary circulation: Pathophysiological relevance of stretch-activated channels in pulmonary hypertension. Biomolecules. 2021;11:1389
  80. 80. Canales COB, Mayor R. Mechanosensitive ion channels in cell migration. Cells and Development. 2021;166:203683. DOI: 10.1016/j.cdev.2021.203683. Epub 2021 Apr 27
  81. 81. Caterina MJ, Rosen TA, Tominaga M, Brake AJ, Julius D. A capsaicin-receptor homologue with a high threshold for noxious heat. Nature. 1999;398:436-441
  82. 82. Qin N, Neeper MP, Liu Y, Hutchinson TL, Lubin ML, Flores CM. TRPV2 is activated by cannabidiol and mediates CGRP release in cultured rat dorsal root ganglion neurons. The Journal of Neuroscience. 2008;28:6231-6238
  83. 83. Lorin C, Vogeli I, Niggli E. Dystrophic cardiomyopathy: Role of TRPV2 channels in stretch-induced cell damage. Cardiovascular Research. 2015;106:153-162
  84. 84. Shibasaki K. Physiological significance of TRPV2 as a mechanosensor, thermosensor and lipid sensor. The Journal of Physiological Sciences. 2016;66:359-365
  85. 85. McGahon MK, Fernández JA, Dash DP, McKee J, Simpson DA, Zholos AV, et al. TRPV2 channels contribute to stretch-activated cation currents and myogenic constriction in retinal arterioles. Investigative Ophthalmology & Visual Science. 2016;57:5637-5647
  86. 86. Gochman A, Tan XF, Bae C, Chen H, Swartz KJ, Jara-Oseguera A. Cannabidiol sensitizes TRPV2 channels to activation by 2-APB. eLife. 2023;12:e86166. DOI: 10.7554/eLife.86166
  87. 87. Conde J, Pumroy RA, Baker C, Rodrigues T, Guerreiro A, Sousa BB, et al. Allosteric antagonist modulation of TRPV2 by Piperlongumine impairs glioblastoma progression. ACS Central Science. 2021;7:868-881
  88. 88. Suzuki M, Mizuno A, Kodaira K, Imai M. Impaired pressure sensation in mice lacking TRPV4. The Journal of Biological Chemistry. 2003;278:22664-22668
  89. 89. Liedtke W, Choe Y, Marti-Renom MA, Bell AM, Denis CS, Sali A, et al. Vanilloid receptor-related osmotically activated channel (VR-OAC), a candidate vertebrate osmoreceptor. Cell. 2000;103:525-535
  90. 90. Alessandri-Haber N, Yeh JJ, Boyd AE, Parada CA, Chen X, Reichling DB, et al. Hypotonicity induces TRPV4-mediated nociception in rat. Neuron. 2003;39:497-511
  91. 91. Mizuno A, Matsumoto N, Imai M, Suzuki M. Impaired osmotic sensation in mice lacking TRPV4. American Journal of Physiology. Cell Physiology. 2003;285:C96-C101
  92. 92. Liedtke W, Tobin DM, Bargmann CI, Friedman JM. Mammalian TRPV4 (VR-OAC) directs behavioral responses to osmotic and mechanical stimuli in Caenorhabditis elegans. Proceedings of the National Academy of Sciences of the United States of America. 2003;100(Suppl. 2):14531-14536. Epub 2003 Oct 27
  93. 93. Guler AD, Lee H, Iida T, Shimizu I, Tominaga M, Caterina M. Heat-evoked activation of the ion channel, TRPV4. The Journal of Neuroscience. 2002;22:6408-6414
  94. 94. Watanabe H, Vriens J, Suh SH, Benham CD, Droogmans G, Nilius B. Heat-evoked activation of TRPV4 channels in a HEK293 cell expression system and in native mouse aorta endothelial cells. The Journal of Biological Chemistry. 2002;277:47044-47051
  95. 95. Chung MK, Lee H, Caterina MJ. Warm temperatures activate TRPV4 in mouse 308 keratinocytes. The Journal of Biological Chemistry. 2003;278:32037-32046
  96. 96. Lee H, Iida T, Mizuno A, Suzuki M, Caterina MJ. Altered thermal selection behavior in mice lacking transient receptor potential vanilloid 4. The Journal of Neuroscience. 2005;25:1304-1310
  97. 97. Watanabe H, Vriens J, Prenen J, Droogmans G, Voets T, Nilius B. Anandamide and arachidonic acid use epoxyeicosatrienoic acids to activate TRPV4 channels. Nature. 2003;424:434-438
  98. 98. Fernández-Carvajal A, Fernández-Ballester G, González-Muñiz R, Ferrer-Montiel A. Chapter 2 pharmacology of TRP channels. In: Madrid R, Bacigalupo J, editors. TRP Channels in Sensory Transduction. Alicante, Spain: Springer International Publishing Switzerland; 2015. pp. 41-50
  99. 99. Lawhorn BG, Brnardic EJ, Behm DJ. Recent advances in TRPV4 agonists and antagonists. Bioorganic & Medicinal Chemistry Letters. 2020;30:127022
  100. 100. Grace MS, Bonvini SJ, Belvisi MG, McIntyre P. Modulation of the TRPV4 ion channel as a therapeutic target for disease. Pharmacology & Therapeutics. 2017;177:9-22
  101. 101. Sappington RM, Sidorova T, Ward NJ, Chakravarthy R, Ho KW, Calkins DJ. Activation of transient receptor potential vanilloid-1 (TRPV1) influences how retinal ganglion cell neurons respond to pressure-related stress. Channels (Austin, Tex.). 2015;9:102-113
  102. 102. Ciura S, Liedtke W, Bourque CW. Hypertonicity sensing in organum vasculosum lamina terminalis neurons: A mechanical process involving TRPV1 but not TRPV4. The Journal of Neuroscience. 2011;31:14669-14676
  103. 103. Lawson JJ, McIlwrath SL, Woodbury CJ, Davis BM, Koerber HR. TRPV1 unlike TRPV2 is restricted to a subset of mechanically insensitive cutaneous nociceptors responding to heat. The Journal of Pain. 2008;9:298-308
  104. 104. Voets T, Droogmans G, Wissenbach U, Janssens A, Flockerzi V, Nilius B. The principle of temperature-dependent gating in cold- and heat-sensitive TRP channels. Nature. 2004;430:748-754
  105. 105. Mugo A, Chou R, Chin F, Liu B, Jiang QX, Qin F. A suicidal mechanism for the exquisite temperature sensitivity of TRPV1. Proceedings of the National Academy of Sciences of the United States of America. 2023;120:e2300305120
  106. 106. Leonelli M, Martins DO, Britto LR. Retinal cell death induced by TRPV1 activation involves NMDA signaling and upregulation of nitric oxide synthases. Cellular and Molecular Neurobiology. 2013;33:379-392
  107. 107. Ho KW, Lambert WS, Calkins DJ. Activation of the TRPV1 cation channel contributes to stress-induced astrocyte migration. Glia. 2014;62:1435-1451
  108. 108. Sakamoto K, Kuroki T, Okuno Y, Sekiya H, Watanabe A, Sagawa T, et al. Activation of the TRPV1 channel attenuates N-methyl-D-aspartic acid-induced neuronal injury in the rat retina. European Journal of Pharmacology. 2014;733:13-22
  109. 109. Pecze L, Winter Z, Josvay K, Otvos F, Kolozsi C, Vizler C, et al. Divalent heavy metal cations block the TRPV1 Ca(2+) channel. Biological Trace Element Research. 2013;151:451-461
  110. 110. Gilliam JC, Wensel TG. TRP channel gene expression in the mouse retina. Vision Research. 2011;51:2440-2452
  111. 111. Yazulla S, Studholme KM. Vanilloid receptor like 1 (VRL1) immunoreactivity in mammalian retina: Colocalization with somatostatin and purinergic P2X1 receptors. The Journal of Comparative Neurology. 2004;474:407-418
  112. 112. Leonelli M, Martins DO, Kihara AH, Britto LR. Ontogenetic expression of the vanilloid receptors TRPV1 and TRPV2 in the rat retina. International Journal of Developmental Neuroscience. 2009;27:709-718
  113. 113. Zimov S, Yazulla S. Localization of vanilloid receptor 1 (TRPV1/VR1)-like immunoreactivity in goldfish and zebrafish retinas: Restriction to photoreceptor synaptic ribbons. Journal of Neurocytology. 2004;33:441-452
  114. 114. Ryskamp DA, Witkovsky P, Barabas P, Huang W, Koehler C, Akimov NP, et al. The polymodal ion channel transient receptor potential vanilloid 4 modulates calcium flux, spiking rate, and apoptosis of mouse retinal ganglion cells. The Journal of Neuroscience. 2011;31:7089-7101
  115. 115. Taylor L, Arner K, Ghosh F. Specific inhibition of TRPV4 enhances retinal ganglion cell survival in adult porcine retinal explants. Experimental Eye Research. 2016;154:10-21
  116. 116. Yarishkin O, Phuong TTT, Lakk M, Krizaj D. TRPV4 does not regulate the distal retinal light response. Advances in Experimental Medicine and Biology. 2018;1074:553-560. DOI: 10.1007/978-3-319-75402-4_67
  117. 117. Jo AO, Lakk M, Rudzitis CN, Krizaj D. TRPV4 and TRPC1 channels mediate the response to tensile strain in mouse Muller cells. Cell Calcium. 2022;104:102588. DOI: 10.1016/j.ceca.2022.102588. Epub 2022 Apr 5
  118. 118. Li Q, Cheng Y, Zhang S, Sun X, Wu J. TRPV4-induced Muller cell gliosis and TNF-a elevation-mediated retinal ganglion cell apoptosis in glaucomatous rats via JAK2/STAT3/NF-kB pathway. Journal of Neuroinflammation. 2021;18:271-02315
  119. 119. Redmon SN, Yarishkin O, Lakk M, Jo A, Mustafic E, Tvrdik P, et al. TRPV4 channels mediate the mechanoresponse in retinal microglia. Glia. 2021;69:1563-1582
  120. 120. Matsumoto H, Sugio S, Seghers F, Krizaj D, Akiyama H, Ishizaki Y, et al. Retinal detachment-induced Muller glial cell swelling activates TRPV4 ion channels and triggers photoreceptor death at body temperature. The Journal of Neuroscience. 2018;38:8745-8758
  121. 121. Nilius B, Vriens J, Prenen J, Droogmans G, Voets T. TRPV4 calcium entry channel: A paradigm for gating diversity. American Journal of Physiology. Cell Physiology. 2004;286:C195-C205
  122. 122. Shibasaki K, Sugio S, Takao K, Yamanaka A, Miyakawa T, Tominaga M, et al. TRPV4 activation at the physiological temperature is a critical determinant of neuronal excitability and behavior. Pflügers Archiv. 2015;467:2495-2507
  123. 123. Sawamura S, Shirakawa H, Nakagawa T, Mori Y. Y. Kaneko TRP channels in the brain: What are they there for? In: Emir TLR, editor. Neurobiology of TRP Channels. Boca Raton (FL): CRC Press/Taylor & Francis; 2017
  124. 124. Ma L, Liu X, Liu Q, Jin S, Chang H, Liu H. The roles of transient receptor potential ion channels in pathologies of glaucoma. Frontiers in Physiology. 2022;13:806786. DOI: 10.3389/fphys.2022.806786. eCollection 2022
  125. 125. Dosey TL, Wang Z, Fan G, Zhang Z, Serysheva II, Chiu W, et al. Structures of TRPV2 in distinct conformations provide insight into role of the pore turret. Nature Structural & Molecular Biology. 2019;26:40-49
  126. 126. Huynh KW, Cohen MR, Jiang J, Samanta A, Lodowski DT, Zhou ZH, et al. Structure of the full-length TRPV2 channel by cryo-EM. Nature Communications. 2016;7:11130
  127. 127. Barro-Soria R, Stindl J, Muller C, Foeckler R, Todorov V, Castrop H, et al. Angiotensin-2-mediated Ca2+ signaling in the retinal pigment epithelium: Role of angiotensin-receptor-associated-protein and TRPV2 channel. PLoS One. 2012;7:e49624
  128. 128. Reichhart N, Keckeis S, Fried F, Fels G, Strauss O. Regulation of surface expression of TRPV2 channels in the retinal pigment epithelium. Graefe's Archive for Clinical and Experimental Ophthalmology. 2015;253:865-874
  129. 129. Cordeiro S, Seyler S, Stindl J, Milenkovic VM, Strauss O. Heat-sensitive TRPV channels in retinal pigment epithelial cells: Regulation of VEGF-A secretion. Investigative Ophthalmology & Visual Science. 2010;51:6001-6008
  130. 130. Zhao PY, Gan G, Peng S, Wang SB, Chen B, Adelman RA, et al. TRP channels localize to subdomains of the apical plasma membrane in human Fetal retinal pigment epithelium. Investigative Ophthalmology & Visual Science. 2015;56:1916-1923
  131. 131. Mirshahi M, Nicolas C, Mirshahi S, Golestaneh N, d'Hermies F, Agarwal MK. Immunochemical analysis of the sodium channel in rodent and human eye. Experimental Eye Research. 1999;69:21-32
  132. 132. Canessa CM, Schild L, Buell G, Thorens B, Gautschi I, Horisberger JD, et al. Amiloride-sensitive epithelial Na+ channel is made of three homologous subunits. Nature. 1994;367:463-467
  133. 133. Lemmens-Gruber R, Tzotzos S. The epithelial Sodium Channel-an underestimated drug target. International Journal of Molecular Sciences. 2023;24:7775
  134. 134. Golestaneh N, Nicolas C, Picaud S, Ferrari P, Mirshahi M. The epithelial sodium channel (ENaC) in rodent retina, ontogeny and molecular identity. Current Eye Research. 2000;21:703-709
  135. 135. Brockway LM, Benos DJ, Keyser KT, Kraft TW. Blockade of amiloride-sensitive sodium channels alters multiple components of the mammalian electroretinogram. Visual Neuroscience. 2005;22:143-151
  136. 136. Aufy M, Hussein AM, Stojanovic T, Studenik CR, Kotob MH. Proteolytic activation of the epithelial Sodium Channel (ENaC): Its mechanisms and implications. International Journal of Molecular Sciences. 2023;24:17563
  137. 137. O'Hagan R, Chalfie M, Goodman MB. The MEC-4 DEG/ENaC channel of Caenorhabditis elegans touch receptor neurons transduces mechanical signals. Nature Neuroscience. 2005;8:43-50
  138. 138. Hamill OP, McBride DW Jr. A supramolecular complex underlying touch sensitivity. Trends in Neurosciences. 1996;19:258-261
  139. 139. Chalfant ML, Denton JS, Berdiev BK, Ismailov II, Benos DJ, Stanton BA. Intracellular H+ regulates the alpha-subunit of ENaC, the epithelial Na+ channel. The American Journal of Physiology. 1999;276:C477-C486
  140. 140. Anantharam A, Tian Y, Palmer LG. Open probability of the epithelial sodium channel is regulated by intracellular sodium. The Journal of Physiology. 2006;574:333-347
  141. 141. Planes C, Caughey GH. Regulation of the epithelial Na+ channel by peptidases. Current Topics in Developmental Biology. 2007;78:23-46. DOI: 10.1016/S0070-2153
  142. 142. Lu M, Echeverri F, Kalabat D, Laita B, Dahan DS, Smith RD, et al. Small molecule activator of the human epithelial sodium channel. The Journal of Biological Chemistry. 2008;283:11981-11994
  143. 143. Butterworth MB, Zhang L, Heidrich EM, Myerburg MM, Thibodeau PH. Activation of the epithelial sodium channel (ENaC) by the alkaline protease from Pseudomonas aeruginosa. The Journal of Biological Chemistry. 2012;287:32556-32565
  144. 144. Anand D, Hummler E, Rickman OJ. ENaC activation by proteases. Acta Physiologica (Oxford, England). 2022;235:e13811
  145. 145. Schild L. The epithelial sodium channel and the control of sodium balance. Biochimica et Biophysica Acta. 1802;2010:1159-1165
  146. 146. Shabbir W, Scherbaum-Hazemi P, Tzotzos S, Fischer B, Fischer H, Pietschmann H, et al. Mechanism of action of novel lung edema therapeutic AP301 by activation of the epithelial sodium channel. Molecular Pharmacology. 2013;84:899-910
  147. 147. Tzotzos S, Fischer B, Fischer H, Pietschmann H, Lucas R, Dupré G, et al. AP301, a synthetic peptide mimicking the lectin-like domain of TNF, enhances amiloride-sensitive Na(+) current in primary dog, pig and rat alveolar type II cells. Pulmonary Pharmacology & Therapeutics. 2013;26:356-363
  148. 148. Schmid B, Kredel M, Ullrich R, Krenn K, Lucas R, Markstaller K, et al. Safety and preliminary efficacy of sequential multiple ascending doses of solnatide to treat pulmonary permeability edema in patients with moderate-to-severe ARDS-a randomized, placebo-controlled, double-blind trial. Trials. 2021;22(1):643
  149. 149. Coste B, Mathur J, Schmidt M, Earley TJ, Ranade S, Petrus MJ, et al. Piezo1 and Piezo2 are essential components of distinct mechanically activated cation channels. Science. 2010;330:55-60
  150. 150. Coste B, Xiao B, Santos JS, Syeda R, Grandl J, Spencer KS, et al. Piezo proteins are pore-forming subunits of mechanically activated channels. Nature. 2012;19(483):176-181
  151. 151. Xiao R, Xu XS. Mechanosensitive channels: In touch with piezo. Current Biology. 2010;20:R936-R938
  152. 152. Gottlieb PA, Sachs F. The sensation of stretch. Nature. 2012;483:163-164
  153. 153. Martins JR, Penton D, Peyronnet R, Arhatte M, Moro C, Picard N, et al. Piezo1-dependent regulation of urinary osmolarity. Pflügers Archiv. 2016;468:1197-1206
  154. 154. Ridone P, Vassalli M, Martinac B. Piezo1 mechanosensitive channels: What are they and why are they important. Biophysical Reviews. 2019;11:795-805
  155. 155. Stewart L, Turner NA. Channelling the force to reprogram the matrix: Mechanosensitive ion channels in cardiac fibroblasts. Cells. 2021;10:990
  156. 156. Gnanasambandam R, Bae C, Gottlieb PA, Sachs F. Ionic selectivity and permeation properties of human PIEZO1 channels. PLoS One. 2015;10:e0125503
  157. 157. Gottlieb PA, Sachs F. Piezo1: Properties of a cation selective mechanical channel. Channels (Austin, Tex.). 2012;6:214-219
  158. 158. Baxter SL, Keenan WT, Athanas AJ, Proudfoot JA, Zangwill LM, Ayyagari R, et al. Investigation of associations between Piezo1 mechanoreceptor gain-of-function variants and glaucoma-related phenotypes in humans and mice. Scientific Reports. 2020;10:19013-76026
  159. 159. Zhu Y, Garcia-Sanchez J, Dalal R, Sun Y, Kapiloff MS, Goldberg JL, et al. Differential expression of PIEZO1 and PIEZO2 mechanosensitive channels in ocular tissues implicates diverse functional roles. Experimental Eye Research;236:2023, 109675. DOI: 10.1016/j.exer.2023.109675. Epub 2023 Oct 10
  160. 160. Bocchero U, Falleroni F, Mortal S, Li Y, Cojoc D, Lamb T, et al. Mechanosensitivity is an essential component of phototransduction in vertebrate rods. PLOS Biology. 2020;18:e3000750
  161. 161. Pang JJ, Yang Z, Jacoby RA, Wu SM. Cone synapses in mammalian retinal rod bipolar cells. The Journal of Comparative Neurology. 2018;526:1896-1909
  162. 162. Sun X, Hirano AA, Brecha NC, Barnes S. Calcium-activated BK(Ca) channels govern dynamic membrane depolarizations of horizontal cells in rodent retina. The Journal of Physiology. 2017;595:4449-4465
  163. 163. Fernandez-Sanchez L, de Sevilla Muller LP, Brecha NC, Cuenca N. Loss of outer retinal neurons and circuitry alterations in the DBA/2J mouse. Investigative Ophthalmology & Visual Science. 2014;55:6059-6072
  164. 164. Bayer AU, Neuhardt T, May AC, Martus P, Maag KP, Brodie S, et al. Retinal morphology and ERG response in the DBA/2NNia mouse model of angle-closure glaucoma. Investigative Ophthalmology & Visual Science. 2001;42:1258-1265
  165. 165. Burrone J, Lagnado L. Electrical resonance and Ca2+ influx in the synaptic terminal of depolarizing bipolar cells from the goldfish retina. The Journal of Physiology. 1997;505:571-584
  166. 166. Tanimoto N, Sothilingam V, Euler T, Ruth P, Seeliger MW, Schubert T. BK channels mediate pathway-specific modulation of visual signals in the in vivo mouse retina. The Journal of Neuroscience. 2012;32:4861-4866
  167. 167. Sausbier M, Hu H, Arntz C, Feil S, Kamm S, Adelsberger H, et al. Cerebellar ataxia and Purkinje cell dysfunction caused by Ca2+−activated K+ channel deficiency. Proceedings of the National Academy of Sciences of the United States of America. 2004;101:9474-9478
  168. 168. Dyka FM, May CA, Enz R. Subunits of the epithelial sodium channel family are differentially expressed in the retina of mice with ocular hypertension. Journal of Neurochemistry. 2005;94:120-128
  169. 169. Crousillac S, LeRouge M, Rankin M, Gleason E. Immunolocalization of TRPC channel subunits 1 and 4 in the chicken retina. Visual Neuroscience. 2003;20:453-463
  170. 170. Krueger B, Schlotzer-Schrehardt U, Haerteis S, Zenkel M, Chankiewitz VE, Amann KU, et al. Four subunits (abgd) of the epithelial sodium channel (ENaC) are expressed in the human eye in various locations. Investigative Ophthalmology & Visual Science. 2012;53:596-604
  171. 171. Reyes R, Lauritzen I, Lesage F, Ettaiche M, Fosset M, Lazdunski M. Immunolocalization of the arachidonic acid and mechanosensitive baseline traak potassium channel in the nervous system. Neuroscience. 2000;95:893-901
  172. 172. Weitlauf C, Ward NJ, Lambert WS, Sidorova TN, Ho KW, Sappington RM, et al. Short-term increases in transient receptor potential vanilloid-1 mediate stress-induced enhancement of neuronal excitation. The Journal of Neuroscience. 2014;34:15369-15381
  173. 173. Noailles A, Kutsyr O, Mayordomo-Febrer A, Lax P, López-Murcia M, Sanz-González SM, et al. Sodium hyaluronate-induced ocular hypertension in rats damages the direction-selective circuit and inner/outer retinal plexiform layers. Investigative Ophthalmology & Visual Science. 2022;63:2
  174. 174. Agostinone J, Di PA. Retinal ganglion cell dendrite pathology and synapse loss: Implications for glaucoma. Progress in Brain Research. 2015;220:199-216
  175. 175. Berry RH, Qu J, John SW, Howell GR, Jakobs TC. Synapse loss and dendrite Remodeling in a mouse model of glaucoma. PLoS One. 2015;10:e0144341
  176. 176. El-Danaf RN, Huberman AD. Characteristic patterns of dendritic remodeling in early-stage glaucoma: Evidence from genetically identified retinal ganglion cell types. The Journal of Neuroscience. 2015;35:2329-2343
  177. 177. DeMar J, Sharrow K, Hill M, Berman J, Oliver T, Long J. Effects of primary blast overpressure on retina and optic tract in rats. Frontiers in Neurology. 2016;7:59. DOI: 10.3389/fneur.2016.00059. eCollection 2016
  178. 178. Zhu Y, Howard JT, Edsall PR, Morris RB, Lund BJ, Cleland JM. Blast exposure induces ocular functional changes with increasing blast over-pressures in a rat model. Current Eye Research. 2019;44:770-780
  179. 179. Naguib S, Bernardo-Colon A, Cencer C, Gandra N, Rex TS. Galantamine protects against synaptic, axonal, and vision deficits in experimental neurotrauma. Neurobiology of Disease. 2020;134:104695. DOI: 10.1016/j.nbd.2019.104695. Epub 2019 Nov 25
  180. 180. Allen RS, Motz CT, Feola A, Chesler KC, Haider R, Ramachandra RS, et al. Long-term functional and structural consequences of primary blast overpressure to the eye. Journal of Neurotrauma. 2018;35:2104-2116
  181. 181. Bricker-Anthony C, Hines-Beard J, Rex TS. Molecular changes and vision loss in a mouse model of closed-globe blast trauma. Investigative Ophthalmology & Visual Science. 2014;55:4853-4862
  182. 182. North RV, Jones AL, Drasdo N, Wild JM, Morgan JE. Electrophysiological evidence of early functional damage in glaucoma and ocular hypertension. Investigative Ophthalmology & Visual Science. 2010;51:1216-1222
  183. 183. Drasdo N, Aldebasi YH, Chiti Z, Mortlock KE, Morgan JE, North RV. The s-cone PHNR and pattern ERG in primary open angle glaucoma. Investigative Ophthalmology & Visual Science. 2001;42:1266-1272
  184. 184. Frankfort BJ, Khan AK, Tse DY, Chung I, Pang JJ, Yang Z, et al. Elevated intraocular pressure causes inner retinal dysfunction before cell loss in a mouse model of experimental glaucoma. Investigative Ophthalmology & Visual Science. 2013;54:762-770
  185. 185. Yang TH, Kang EY, Lin PH, Wu PL, Sachs JA, Wang NK. The value of electroretinography in identifying candidate genes for inherited retinal dystrophies: A diagnostic guide. Diagnostics (Basel). 2023;13:3041
  186. 186. Kurtenbach A, Kramer S, Strasser T, Zrenner E, Langrová H. The importance of electrode position in visual electrophysiology. Documenta Ophthalmologica. 2017;134:129-134
  187. 187. Velten IM, Korth M, Horn FK. The a-wave of the dark adapted electroretinogram in glaucomas: Are photoreceptors affected? The British Journal of Ophthalmology. 2001;85:397-402
  188. 188. Cuenca N, Pinilla I, Fernandez-Sanchez L, Salinas-Navarro M, Alarcon-Martinez L, Aviles-Trigueros M, et al. Changes in the inner and outer retinal layers after acute increase of the intraocular pressure in adult albino Swiss mice. Experimental Eye Research. 2010;91:273-285
  189. 189. Velten IM, Horn FK, Korth M, Velten K. The b-wave of the dark adapted flash electroretinogram in patients with advanced asymmetrical glaucoma and normal subjects. The British Journal of Ophthalmology. 2001;85:403-409
  190. 190. Shen Y, Luo X, Liu S, Shen Y, Nawy S, Shen Y. Rod bipolar cells dysfunction occurs before ganglion cells loss in excitotoxin-damaged mouse retina. Cell Death & Disease. 2019;10:905-2140
  191. 191. Mitra P, Slaughter MM. Mechanism of generation of spontaneous miniature outward currents (SMOCs) in retinal amacrine cells. The Journal of General Physiology. 2002;119:355-372
  192. 192. Grimes WN, Li W, Chávez AE, Diamond JS. BK channels modulate pre- and postsynaptic signaling at reciprocal synapses in retina. Nature Neuroscience. 2009;12:585-592
  193. 193. Oda M, Yamamoto H, Matsumoto H, Ishizaki Y, Shibasaki K. TRPC5 regulates axonal outgrowth in developing retinal ganglion cells. Laboratory Investigation. 2020;100:297-310
  194. 194. Ford KJ, Arroyo DA, Kay JN, Lloyd EE, Bryan RM Jr, Sanes JR, et al. A role for TREK1 in generating the slow afterhyperpolarization in developing starburst amacrine cells. Journal of Neurophysiology. 2013;109:2250-2259
  195. 195. Li Q, Jin R, Zhang S, Sun X, Wu J. Transient receptor potential vanilloid four channels modulate inhibitory inputs through differential regulation of GABA and glycine receptors in rat retinal ganglion cells. The FASEB Journal. 2020;34:14521-14538
  196. 196. Buchi ER. Cell death in the rat retina after a pressure-induced ischaemia-reperfusion insult: An electron microscopic study. I. Ganglion cell layer and inner nuclear layer. Experimental Eye Research. 1992;55:605-613
  197. 197. Pang JJ, Wu SM. Retinal ganglion cell death is correlated with eyeball expansion in mammals. Jacobs Journal of Ophthalmology. 2014;1:1-5
  198. 198. Moon JI, Kim IB, Gwon JS, Park MH, Kang TH, Lim EJ, et al. Changes in retinal neuronal populations in the DBA/2J mouse. Cell and Tissue Research. 2005;320:51-59
  199. 199. Kielczewski JL, Pease ME, Quigley HA. The effect of experimental glaucoma and optic nerve transection on amacrine cells in the rat retina. Investigative Ophthalmology & Visual Science. 2005;46:3188-3196
  200. 200. Pang JJ, Frankfort BJ, Gross RL, Wu SM. Elevated intraocular pressure decreases response sensitivity of inner retinal neurons in experimental glaucoma mice. Proceedings of the National Academy of Sciences of the United States of America. 2015;112:2593-2598
  201. 201. Park HY, Kim JH, Park CK. Alterations of the synapse of the inner retinal layers after chronic intraocular pressure elevation in glaucoma animal model. Molecular Brain. 2014;7:53
  202. 202. Strettoi E, Masland RH. The number of unidentified amacrine cells in the mammalian retina. Proceedings of the National Academy of Sciences of the United States of America. 1996;93:14906-14911
  203. 203. Hernandez M, Rodriguez FD, Sharma SC, Vecino E. Immunohistochemical changes in rat retinas at various time periods of elevated intraocular pressure. Molecular Vision. 2009;15:2696-2709
  204. 204. Akopian A, Kumar S, Ramakrishnan H, Viswanathan S, Bloomfield SA. Amacrine cells coupled to ganglion cells via gap junctions are highly vulnerable in glaucomatous mouse retinas. The Journal of Comparative Neurology. 2019;527:159-173
  205. 205. Moreno MC, De Zavalia N, Sande P, Jaliffa CO, Fernandez DC, Sarmiento MIK, et al. Effect of ocular hypertension on retinal GABAergic activity. Neurochemistry International. 2008;52:675-682
  206. 206. Nemargut JP, Zhu J, Savoie BT, Wang GY. Differential effects of charybdotoxin on the activity of retinal ganglion cells in the dark- and light-adapted mouse retina. Vision Research. 2009;49:388-397
  207. 207. Wang GY, Robinson DW, Chalupa LM. Calcium-activated potassium conductances in retinal ganglion cells of the ferret. Journal of Neurophysiology. 1998;79:151-158
  208. 208. Henne J, Jeserich G. Maturation of spiking activity in trout retinal ganglion cells coincides with upregulation of Kv3.1- and BK-related potassium channels. Journal of Neuroscience Research. 2004;75:44-54
  209. 209. Garaycochea J, Slaughter MM. GABAB receptors enhance excitatory responses in isolated rat retinal ganglion cells. The Journal of Physiology. 2016;594:5543-5554
  210. 210. Zhang XT, Xu Z, Shi KP, Guo DL, Li H, Wang L, et al. Elevated expression of TREK-TRAAK K(2P) channels in the retina of adult rd1 mice. International Journal of Ophthalmology. 2019;12:924-929
  211. 211. Wen X, Liao P, Luo Y, Yang L, Yang H, Liu L, et al. Tandem pore domain acid-sensitive K channel 3 (TASK-3) regulates visual sensitivity in healthy and aging retina. Science Advances. 2022;8:eabn 8785
  212. 212. Choi HJ, Sun D, Jakobs TC. Astrocytes in the optic nerve head express putative mechanosensitive channels. Molecular Vision. 2015;21:749-766. eCollection 2015
  213. 213. Morozumi W, Inagaki S, Iwata Y, Nakamura S, Hara H, Shimazawa M. Piezo channel plays a part in retinal ganglion cell damage. Experimental Eye Research. 2020;191:107900. DOI: 10.1016/j.exer.2019.107900. Epub 2019 Dec 23
  214. 214. Harraz OF, Klug NR, Senatore AJ, Hill-Eubanks DC, Nelson MT. Piezo1 is a Mechanosensor channel in central nervous system capillaries. Circulation Research. 2022;130:1531-1546
  215. 215. Wan Y, Wang H, Fan X, Bao J, Wu S, Liu Q, et al. Mechanosensitive channel Piezo1 is an essential regulator in cell cycle progression of optic nerve head astrocytes. Glia. 2023;71:1233-1246
  216. 216. Zhong W, Lan C, Gu Z, Tan Q, Xiang X, Zhou H, et al. The mechanosensitive Piezo1 channel mediates Mechanochemical transmission in myopic eyes. Investigative Ophthalmology & Visual Science. 2023;64:1
  217. 217. Tang J, Liu Z, Han J, Xue J, Liu L, Lin J, et al. Increased Mobile zinc regulates retinal ganglion cell survival via activating mitochondrial OMA1 and integrated stress response. Antioxidants (Basel). 2022;11:2001
  218. 218. Kim JS, Park JY, Kang HW, Lee EJ, Bang H, Lee JH. Zinc activates TREK-2 potassium channel activity. The Journal of Pharmacology and Experimental Therapeutics. 2005;314:618-625
  219. 219. Lambert S, Drews A, Rizun O, Wagner TF, Lis A, Mannebach S, et al. Transient receptor potential melastatin 1 (TRPM1) is an ion-conducting plasma membrane channel inhibited by zinc ions. The Journal of Biological Chemistry. 2011;286:12221-12233
  220. 220. Jo AO, Ryskamp DA, Phuong TT, Verkman AS, Yarishkin O, MacAulay N, et al. TRPV4 and AQP4 channels synergistically regulate cell volume and calcium homeostasis in retinal Muller glia. The Journal of Neuroscience. 2015;35:13525-13537
  221. 221. Evans LP, Roghair AM, Gilkes NJ, Bassuk AG. Visual outcomes in experimental rodent models of blast-mediated traumatic brain injury. Frontiers in Molecular Neuroscience. 2021;14:659576. DOI: 10.3389/fnmol.2021.659576. eCollection 2021
  222. 222. Mohan K, Kecova H, Hernandez-Merino E, Kardon RH, Harper MM. Retinal ganglion cell damage in an experimental rodent model of blast-mediated traumatic brain injury. Investigative Ophthalmology & Visual Science. 2013;54:3440-3450
  223. 223. Vrabec JP, Levin LA. The neurobiology of cell death in glaucoma. Eye. 2007;21(Suppl. 1):S11-S14
  224. 224. Ou Y, Jo RE, Ullian EM, Wong RO, Della SL. Selective vulnerability of specific retinal ganglion cell types and synapses after transient ocular hypertension. The Journal of Neuroscience. 2016;36:9240-9252
  225. 225. Lakk M, Young D, Baumann JM, Jo AO, Hu H, Krizaj D. Polymodal TRPV1 and TRPV4 sensors Colocalize but do not functionally interact in a subpopulation of mouse retinal ganglion cells. Frontiers in Cellular Neuroscience. 2018;12:353. DOI: 10.3389/fncel.2018.00353. eCollection 2018
  226. 226. Li L, Liu C, Chen L, Chen L. Hypotonicity modulates tetrodotoxin-sensitive sodium current in trigeminal ganglion neurons. Molecular Pain. 2011;7:27. DOI: 10.1186/1744-8069-7-27
  227. 227. Ryskamp DA, Jo AO, Frye AM, Vazquez-Chona F, MacAulay N, Thoreson WB, et al. Swelling and eicosanoid metabolites differentially gate TRPV4 channels in retinal neurons and glia. The Journal of Neuroscience. 2014;19(34):15689-15700
  228. 228. Netti V, Fernandez J, Kalstein M, Pizzoni A, Di GG, Rivarola V, et al. TRPV4 contributes to resting membrane potential in retinal Muller cells: Implications in cell volume regulation. Journal of Cellular Biochemistry. 2017;118:2302-2313
  229. 229. Toft-Bertelsen TL, Yarishkin O, Redmon S, Phuong TTT, Krizaj D, MacAulay N. Volume sensing in the transient receptor potential vanilloid 4 ion channel is cell type-specific and mediated by an N-terminal volume-sensing domain. The Journal of Biological Chemistry. 2019;294:18421-18434
  230. 230. Orduna RM, Noguez IR, Hernandez Godinez NM, Bautista Cortes AM, Lopez Escalante DD, Liedtke W, et al. TRPV4 inhibition prevents increased water diffusion and blood-retina barrier breakdown in the retina of streptozotocin-induced diabetic mice. PLoS One. 2019;14:e0212158
  231. 231. O'Hare M, Esquiva G, McGahon MK, Hombrebueno JMR, Augustine J, Canning P, et al. Loss of TRPV2-mediated blood flow autoregulation recapitulates diabetic retinopathy in rats. JCI Insight. 2022;7:e155128
  232. 232. Nishinaka A, Tanaka M, Ohara K, Sugaru E, Shishido Y, Sugiura A, et al. TRPV4 channels promote vascular permeability in retinal vascular disease. Experimental Eye Research. 2023;228:109405. DOI: 10.1016/j.exer.2023.109405. Epub 2023 Feb 9
  233. 233. Phuong TTT, Redmon SN, Yarishkin O, Winter JM, Li DY, Krizaj D. Calcium influx through TRPV4 channels modulates the adherens contacts between retinal microvascular endothelial cells. The Journal of Physiology. 2017;595:6869-6885
  234. 234. O'Leary C, McGahon MK, Ashraf S, McNaughten J, Friedel T, Cincolà P, et al. Involvement of TRPV1 and TRPV4 channels in retinal angiogenesis. Investigative Ophthalmology & Visual Science. 2019;60:3297-3309
  235. 235. Leonelli M, Martins DO, Britto LR. TRPV1 receptors are involved in protein nitration and Muller cell reaction in the acutely axotomized rat retina. Experimental Eye Research. 2010;91:755-768
  236. 236. Risner ML, McGrady NR, Boal AM, Pasini S, Calkins DJ. TRPV1 supports Axogenic enhanced excitability in response to neurodegenerative stress. Frontiers in Cellular Neuroscience. 2021;14:603419. DOI: 10.3389/fncel.2020.603419. eCollection 2020
  237. 237. Sappington RM, Sidorova T, Long DJ, Calkins DJ. TRPV1: Contribution to retinal ganglion cell apoptosis and increased intracellular Ca2+ with exposure to hydrostatic pressure. Investigative Ophthalmology & Visual Science. 2009;50:717-728
  238. 238. Ward NJ, Ho KW, Lambert WS, Weitlauf C, Calkins DJ. Absence of transient receptor potential vanilloid-1 accelerates stress-induced axonopathy in the optic projection. The Journal of Neuroscience. 2014;34:3161-3170
  239. 239. Rountree CM, Meng C, Troy JB, Saggere L. Mechanical stimulation of the retina: Therapeutic feasibility and cellular mechanism. IEEE Transactions on Neural Systems and Rehabilitation Engineering. 2018;26:1075-1083
  240. 240. Birkholz TR, Beane WS. The planarian TRPA1 homolog mediates extraocular behavioral responses to near-ultraviolet light. The Journal of Experimental Biology. 2017;220:2616-2625
  241. 241. Shen Y, Rampino MA, Carroll RC, Nawy S. G-protein-mediated inhibition of the Trp channel TRPM1 requires the Gbg dimer. Proceedings of the National Academy of Sciences of the United States of America. 2012;109:8752-8757
  242. 242. Xu Y, Orlandi C, Cao Y, Yang S, Choi CI, Pagadala V, et al. The TRPM1 channel in ON-bipolar cells is gated by both the a and the bg subunits of the G-protein go. Scientific Reports. 2016;6:20940. DOI: 10.1038/srep20940
  243. 243. Jimenez I, Prado Y, Marchant F, Otero C, Eltit F, Cabello-Verrugio C, et al. TRPM channels in human diseases. Cells. 2020;9:2604
  244. 244. Contreras E, Nobleman AP, Robinson PR, Schmidt TM. Melanopsin phototransduction: Beyond canonical cascades. The Journal of Experimental Biology. 2021;224:jeb226522
  245. 245. Sachdeva R, Schlotterer A, Schumacher D, Matka C, Mathar I, Dietrich N, et al. TRPC proteins contribute to development of diabetic retinopathy and regulate glyoxalase 1 activity and methylglyoxal accumulation. Molecular Metabolism. 2018;9:156-167. DOI: 10.1016/j.molmet.2018.01.003. Epub 2018 Jan 5
  246. 246. Brown RL, Xiong WH, Peters JH, Tekmen-Clark M, Strycharska-Orczyk I, Reed BT, et al. TRPM3 expression in mouse retina. PLoS One. 2015;10:e0117615
  247. 247. Webster CM, Tworig J, Caval-Holme F, Morgans CW, Feller MB. The impact of steroid activation of TRPM3 on spontaneous activity in the developing retina. eNeuro. 2020;7:ENEURO-19
  248. 248. Martinez-Gil N, Kutsyr O, Fernandez-Sanchez L, Sanchez-Saez X, Albertos-Arranz H, Sanchez-Castillo C, et al. Ischemia-reperfusion increases TRPM7 expression in mouse retinas. International Journal of Molecular Sciences. 2023;24:16068
  249. 249. Hu L, Xu G. Potential protective role of TRPM7 and involvement of PKC/ERK pathway in blue light-induced apoptosis in retinal pigment epithelium cells in vitro. Asia-Pacific Journal of Ophthalmology (Philadelphia, Pa.). 2021;10:572-578

Written By

Ji-Jie Pang

Submitted: 21 February 2024 Reviewed: 03 April 2024 Published: 06 May 2024