Next Article in Journal
Formation of Transient Protein Aggregate-like Centers Is a General Strategy Postponing Degradation of Misfolded Intermediates
Next Article in Special Issue
Anti-Fibrotic and Anti-Inflammatory Role of NO-Sensitive Guanylyl Cyclase in Murine Lung
Previous Article in Journal
Telling Your Right Hand from Your Left: The Effects of DNA Supercoil Handedness on the Actions of Type II Topoisomerases
Previous Article in Special Issue
Characterizing the Protein Isoforms of foraging (for), the PKGI Ortholog in Drosophila melanogaster
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

cGMP Signaling in Photoreceptor Degeneration

Department of Cell Biology, College of Medicine, University of Oklahoma Health Sciences Center, Oklahoma City, OK 73104, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(13), 11200; https://doi.org/10.3390/ijms241311200
Submission received: 23 May 2023 / Revised: 4 July 2023 / Accepted: 5 July 2023 / Published: 7 July 2023

Abstract

:
Photoreceptors in the retina are highly specialized neurons with photosensitive molecules in the outer segment that transform light into chemical and electrical signals, and these signals are ultimately relayed to the visual cortex in the brain to form vision. Photoreceptors are composed of rods and cones. Rods are responsible for dim light vision, whereas cones are responsible for bright light, color vision, and visual acuity. Photoreceptors undergo progressive degeneration over time in many hereditary and age-related retinal diseases. Despite the remarkable heterogeneity of disease-causing genes, environmental factors, and pathogenesis, the progressive death of rod and cone photoreceptors ultimately leads to loss of vision/blindness. There are currently no treatments available for retinal degeneration. Cyclic guanosine 3′, 5′-monophosphate (cGMP) plays a pivotal role in phototransduction. cGMP governs the cyclic nucleotide-gated (CNG) channels on the plasma membrane of the photoreceptor outer segments, thereby regulating membrane potential and signal transmission. By gating the CNG channels, cGMP regulates cellular Ca2+ homeostasis and signal transduction. As a second messenger, cGMP activates the cGMP-dependent protein kinase G (PKG), which regulates numerous targets/cellular events. The dysregulation of cGMP signaling is observed in varieties of photoreceptor/retinal degenerative diseases. Abnormally elevated cGMP signaling interferes with various cellular events, which ultimately leads to photoreceptor degeneration. In line with this, strategies to reduce cellular cGMP signaling result in photoreceptor protection in mouse models of retinal degeneration. The potential mechanisms underlying cGMP signaling-induced photoreceptor degeneration involve the activation of PKG and impaired Ca2+ homeostasis/Ca2+ overload, resulting from overactivation of the CNG channels, as well as the subsequent activation of the downstream cellular stress/death pathways. Thus, targeting the cellular cGMP/PKG signaling and the Ca2+-regulating pathways represents a significant strategy for photoreceptor protection in retinal degenerative diseases.

1. Introduction

The human retina is comprised of light-responsive neuronal cells known as photoreceptors. Light stimulation initiates phototransduction in photoreceptors, which is conveyed to secondary neurons, ganglion cells, and, ultimately, to the visual cortex in the brain to form vision. There are two types of photoreceptors: rods and cones. Rods respond to dim light and enable vision at night, whereas cones react to bright light and are responsible for daylight vision, color vision, and visual acuity [1,2,3,4]. Photoreceptors degenerate in many hereditary retinal diseases, including retinitis pigmentosa (RP: a progressive degeneration of rods leading to blindness), Leber congenital amaurosis (LCA: many LCA forms involve severe dystrophy of photoreceptors), achromatopsia (rod monochromatism, color blindness), cone-rod dystrophy (CRD), and age-related retinal degeneration, such as age-related macular degeneration (AMD). Despite the remarkable heterogeneity of disease-causing genes, environmental factors, and pathogenesis, the progressive death of rod and cone photoreceptors ultimately leads to the loss of vision/blindness.
Cyclic guanosine 3′, 5′-monophosphate (cGMP) acts as a second messenger molecule in almost all cell types by activating the cGMP-dependent protein kinase G (PKG) and regulating various cellular activities. In photoreceptors, besides acting as a second messenger, cGMP plays a pivotal role in phototransduction by governing the cyclic nucleotide-gated (CNG) channels. In darkness, CNG channels are kept open by the binding of cGMP, maintaining a steady Ca2+ and Na+ influx. Upon light stimulation, the light-sensing protein rhodopsin/opsin undergoes a conformational rearrangement, which activates the heterotrimeric G protein transducin. Consequently, the α subunit of transducin activates phosphodiesterase 6 (PDE6), leading to the catalytic acceleration of cGMP hydrolysis, closure of the channels, and the hyperpolarization of the cell [5,6,7,8,9] (Figure 1). Phototransduction follows the same cascade in both rods and cones, even though the components involved in the cascade vary in the two types of the photoreceptors. For example, the rod CNG channel is composed of CNGA1 and CNGB1 subunits, whereas the cone CNG channel is composed of CNGA3 and CNGB3 [6,10]. The rod PDE6 is composed of PDE6α, PDE6β, and PDE6γ subunits, whereas the cone PDE6 is composed of the PDE6α’ and PDE6γ’ subunits [11,12].
Among the various cellular alterations that occur in photoreceptor degeneration and the potential causative factors, dysregulation of cGMP signaling appears to be one of the leading features. The dysregulation of cGMP signaling is associated with the pathogenesis of RP, LCA, achromatopsia, and CRD [14], and manipulations that reduce the cellular cGMP level result in photoreceptor protection in animal models of retinal degeneration. Here, we review the regulation of cGMP signaling in photoreceptors, the dysregulation of cGMP signaling in photoreceptor degeneration, the experimental evidence supporting the contribution of the elevated cGMP signaling to photoreceptor degeneration, and the potential mechanisms underlying cGMP signaling-induced photoreceptor degeneration. A better understanding of cGMP signaling in photoreceptor degeneration will help the development of therapeutic interventions targeting this signaling pathway for photoreceptor protection.

2. Regulation of Cellular cGMP Level in Photoreceptors

The cellular level of cGMP is highly regulated in photoreceptors. The biosynthesis of cGMP is catalyzed by the retinal guanylyl cyclases (RetGCs) from the guanosine triphosphate (GTP), and its degradation is achieved by PDE6, which specifically hydrolyzes the 3′5′ cyclic phosphate bond [15] (Figure 1). The biosynthesis of cGMP is tightly regulated by guanylate cyclase activator proteins (GCAPs) [16,17]. There are four basic elements regulating cellular cGMP levels in photoreceptors, e.g., PDE6, RetGC, GCAP, and intracellular Ca2+ level ([Ca2+]i). In addition, the functionality of the photoreceptor CNG channels affects the cellular cGMP levels indirectly via [Ca2+]i.

2.1. The Ca2+/Mg2+-GCAP/RetGC Complex

The production of cGMP in photoreceptors is catalyzed by RetGCs. There are two subtypes of RetGCs: RetGC1 and RetGC2. RetGC1 is encoded by GUCY2D (Gucy2e in mouse), and RetGC2 is encoded by GUCY2F (Gucy2f in mouse) [16,18]. RetGC1 is the main isozyme of the cyclases, expressed in both rods and cones, whereas RetGC2 is primarily expressed in rods, and it is nearly undetectable in cones [19,20]. RetGC1 and RetGC2 act as homodimers and share similar domain structures with other membrane-bound guanylyl cyclases [21,22,23]. They both contain an extracellular ligand-binding domain, a dimerization domain, intracellular protein kinase homology domains, and cyclase catalytic domains. Although the three-dimensional structure of RetGC has not been determined, its working pattern has been demonstrated. The two opposing subunits of the RetGC homodimer bind to two Mg2+-GTP substrate molecules, simultaneously with their catalytic domains, creating a single active site to convert GTP to cGMP [24,25]. RetGC1 is responsible for over 70% of the cGMP production in photoreceptors, whereas the RetGC2 isozyme serves as the ancillary component, producing less than 30 percent of the total cGMP [20].
The activity of RetGC is highly regulated by GCAP which binds to RetGC to form a RetGC/GCAP complex (Figure 1). There are two subtypes of GCAPs: GCAP1 and GCAP2. GCAP1 virtually regulates RetGC1 only, whereas GCAP2 activates both RetGC1 and RetGC2 [26,27,28]. The structure of the RetGC/GCAP complex remains less clear. It is known that the activity of RetGC, in the complex, is correlated strongly with the Ca2+ or Mg2+ ligands of GCAP. GCAPs are recoverin-like Ca2+ sensor proteins containing a helix–loop–helix metal-binding structure [29,30]. When the helix–loop–helix metal-binding domains are occupied by Ca2+, the complex inhibits RetGC and reduces cGMP synthesis. When the metal-binding domains are occupied by Mg2+, the complex activates RetGC and stimulates cGMP synthesis [17,31,32]. In dark-adapted photoreceptors, where [Ca2+]i is high, the Ca2+-bound GCAPs inactivate RetGC and inhibit cGMP synthesis (Figure 1). In light-activated photoreceptors, where [Ca2+]i is low, the metal-binding domains of GCAPs are replaced by Mg2+ (Mg2+-bound/Ca2+-free form), and the Mg2+-bound GCAPs activate RetGCs and stimulate cGMP synthesis (Figure 1).
In addition, the activity of RetGCs is regulated by the retinal degeneration protein 3 (RD3). RD3 is a 23-kDa monomer with an elongated four-helix bundle [33]. The in vitro and in vivo experiments show that the main role of RD3 is to suppress the aberrant activation of RetGC by GCAPs in the inner segment [34]. It is required for the stability and ciliary trafficking of RetGCs, and it suppresses the activity of the enzymes in the inner segment before trafficking to the outer segment [35].

2.2. PDE6

PDE6 is the primary regulator of intracellular cGMP concentration in rod and cone photoreceptors (Figure 1). PDE enzymes are a superfamily containing 11 members (PDE1-11) in humans, which are structurally related but consist of multiple genes and function distinctly [36]. PDE6 belongs to 3′,5′-cyclic nucleotide PDEs that catalyze the hydrolysis of the phosphodiester bond of cyclic nucleotides [principally, cGMP and cyclic adenosine monophosphate (cAMP)] [37,38]. There are rod PDE6 complexes and cone PDE6 complexes. Rod PDE6 consists of two catalytic subunits—PDE6α and PDE6β—and two identical inhibitory PDE6γ subunits, whereas cone PDE6 is composed of two identical PDE6α’ catalytic subunits and two identical cone-specific PDE6γ’ inhibitory subunits [15,39]. In dark-adapted photoreceptors, the activity of the catalytic subunits is inhibited by the inhibitory γ-subunit at the entrance of the active site [40]. Upon light stimulation, the photopigment rhodopsin/opsins undergo a conformational change, which activates transducin, and the α subunit of transducin binds to the γ -subunit of PDE6, relieving the inhibitory constraint and activating the enzyme (Figure 1).

2.3. CNG Channel

In addition to regulation by the Ca2+/Mg2+-GCAP/RetGC complex and PDE6, cellular cGMP level is indirectly controlled by photoreceptor CNG channels via [Ca2+]i (Figure 1). Photoreceptor CNG channels are heterotetrameric complexes located on the plasma membranes of the outer segments. They belong to the superfamily of pore–loop cation channels [6,10]. The CNG channel family is comprised of six homologous members, which are classified as structurally similar A subunits (CNGA1-4) and B subunits (CNGB1 and CNGB3), assembled into distinct cell types/specific combinations of heterotetramers [41]. The rod CNG channel is a heterotetramer consisting of three CNGA1 and one CNGB1 [42,43]. Similarly, the cone CNG channel is a heterotetramer consisting of three CNGA3 and one CNGB3 [44,45]. CNG channels are strictly ligand-gated ion channels because the channel-opening state requires the binding of cGMP or cAMP to their cyclic nucleotide-binding domain [46]. Opening the channels allows for an influx of Na+ and Ca2+ ions. In darkness, Ca2+ ions carry about 20% of the ionic current flowing through the channel in rods, whereas this fraction is about 35% in cones [47]. The resulting influx of Na+ and Ca2+ (so called “dark current”) keeps the plasma membrane depolarized and promotes the synaptic glutamate release. As a feedback regulatory mechanism, the elevated [Ca2+]i inhibits cGMP synthesis via the GCAP/RetGC complex. The elevated [Ca2+]i also slightly reduces the affinity of cGMP to the channel [48]. Light triggers hydrolysis of cGMP, leading to the closure of CNG channels, reduction in Ca2+ influx/[Ca2+]i, and the subsequent activation of the GCAP/RetGC complex (Figure 1). Thus, CNG channels not only play a pivotal role in phototransduction but also regulate cellular cGMP levels.

3. Dysregulation of cGMP Signaling in Photoreceptor Degeneration

Among the various cellular alterations that occur in photoreceptor degeneration and the potential causative factors, dysregulation of cGMP signaling represents one of the top pathogeneses. The dysregulation is primarily manifested as the abnormally elevated cellular cGMP levels and the subsequent elevation of PKG signaling, which alters a wide range of cellular activities, resulting in impaired cellular homeostasis, cellular stress, and cell death [14,49]. The accumulation of cGMP also leads to an increase in Ca2+ influx/Ca2+ overload via CNG channels, as well as the subsequent harmful outcomes. Normally, only a small fraction of the total CNG channels is open by cGMP in dark-adapted photoreceptors, and the elevated cGMP leads to opening an excessive number of the channels in the dark, thus abnormally increasing the influx of Ca2+ and Na+. The defects/mutations in various molecules involved in the regulation of cGMP production/degradation, including RetGC1, RD3, GCAP1, PDE6, and CNG channels, are the primary causes of the dysregulation of cGMP signaling in photoreceptors.

3.1. Dysregulation of RetGC1, RD3, and GCAP1

There are over 100 mutations in GUCY2D (encoding RetGC1) identified to date. These mutations are mainly associated with autosomal-dominant CRD (adCRD) [50,51,52,53,54,55] and autosomal-recessive LCA (arLCA) [56,57,58,59,60]. The dominant, gain-of-function mutations in GUCY2D are associated with the degeneration of both rods and cones, causing adCRD, whereas the GUCY2D LCA mutation (arLCA) primarily results in the dysfunction of signaling, leading to the loss of photoreceptor function/blindness [61]. Mutations in GUCY2D account for about 40% of the total adCRD cases and 10–20% of the total arLCA cases [53]. The adCRD mutations are all missense variants, and the mutant proteins are functional. Studies using animal models and cell culture models show that many of these mutations enhance the sensitivity of RetGC1 to GCAP, resulting in excessive cGMP production and enhanced Ca2+ influx [55,62,63,64]. The residue Arg838, located in the dimerization domain of RetGC1, is a mutation hotspot for adCRD, and it is the typical example of overactivation dysregulation. The phenotype in R838S RetGC1 transgenic mice is completely prevented by the deletion of GCAP [63]. In contrast, the LCA phenotype is associated with biallelic null mutations in which the RetGC1 activity is severely reduced or absent due to impaired sensitivity to GCAPs [54,56,57], thereby leading to less production of cGMP and dysfunction of signaling. Of note, the GUCY2D LCA mutations do not necessarily result in the degeneration of photoreceptors. Although GUCY2D arLCA patients display retinal blindness from birth, the vast majority of photoreceptors remain alive throughout adulthood. It was recently shown that GUCY2D replacement by gene therapy quickly improves cone and rod sensitivity after decades of blindness [65,66]. Other, less common phenotypes associated with mutations in GUCY2D include autosomal recessive CRD (arCRD) [67] and autosomal recessive congenital stationary night blindness (arCSNB) [68]. It should be noted that the GUCY2D arCSNB patients show some age-related rod loss in the peripheral retina [68]. In addition, the deficiency of RD3 causes severe photoreceptor degeneration and blindness in recessive LCA, as well as in the rd3 mouse strain [69,70]. This is because a deficiency of the RD3 protein causes the aberrant activation of cGMP production, via the RetGC/GCAP complex in the inner segment, triggering the cell death process.
There are about 20 mutations in GUCA1A (encoding GCAP1) identified to date. These mutations are mainly associated with cone degenerations, including cone dystrophy, cone–rod dystrophy, and macular dystrophy [71,72,73,74,75,76,77]. Studies using animal models and cell culture models show that these mutations, such as the Y99C and G86R mutations [78,79], reduce the sensitivity of GCAP to Ca2+ and subsequently increase RetGC1 activity, leading to the overproduction of cGMP, enhanced CNG channel activity, and enhanced Ca2+ influx [71,72,74,78,79].

3.2. Deficiency of PDE6

Mutations in PDE6A and PDE6B are associated with RP and adCSNB [80,81,82,83,84], whereas mutations in PDE6C and PDE6H are associated with cone dystrophy and achromatopsia [85,86,87,88]. These mutations are all loss-of-function mutations, leading to the cellular accumulation of cGMP and increased Ca2+ influx in photoreceptors. Mouse models with defects of Pde6 genes have been widely studied to understand the disease pathogenesis. The rd1 mouse has a non-sense mutation in the cGMP-binding domain of Pde6b, and as a result, the mRNA/protein is not produced. These mice do not show detectable PDE6 activity [89,90], and they exhibit dramatic increases in cellular cGMP levels between postnatal day 10 (P10) and P14 [91]. The rd1 mice display disorganized outer segments at P7 and the complete loss of rods by about P21, followed by secondary cone degeneration [92,93]. The rd10 mouse harbors an autosomal missense R560C mutation in Pde6b, proceeding with slower rod degeneration compared to rd1 mouse. The Rd10 retinas appear nearly normal for the first 2 weeks, but then, they rapidly deteriorate with a severe loss of photoreceptors [94,95,96]. The mutation does not alter the production of Pde6b mRNA, but it dramatically reduces maximal/basal activity of PDE6, leading to the accumulation of cellular cGMP [97]. The Cpfl1 [cone photoreceptor function loss (1)] mouse harbors a 116-bp insertion between exon 4 and 5, as well as an additional 1-bp deletion in exon 7 in the Pde6c gene, leading to an in-frame shift, introducing premature termination codons, the loss of activity of the enzyme, and the accumulation of cGMP [49,88,98]. Cpfl1 mice display the early-onset, rapid/severe cone degeneration phenotype, including the loss of cone function and death of cones [88,98,99,100].
As mentioned above, the disease-causing mutations in Pde6 genes are all loss-of-function mutations, leading to the cellular accumulation of cGMP and increased Ca2+ influx in photoreceptors. Of note, it seems that the cGMP production feedback via the GCAP/RetGC complex, subsequent to increased [Ca2+]i, cannot decrease the intracellular levels of cGMP in Pde6 mutants. The cellular mechanism(s) underlying this observation is unknown at this time. The observation may suggest that the feedback regulation by the GCAP/RetGC complex is less sensitive to the elevation of [Ca2+]i than to the reduction in [Ca2+]i. It may also imply that the feedback regulation is blunted, for some reason, in the absence of functional PDE6. Nevertheless, the feedback regulation of cGMP production by the GCAP/RetGC complex, in the absence of functional PDE6, merits further investigation.

3.3. Deficiency of CNG Channel

Mutations in CNG channels are associated with photoreceptor degeneration. Mutations in CNGA1 and CNGB1 are associated with arRP, accounting for 2–3% of arRP cases [101,102,103,104]. To date, there are about 40 mutations identified in CNGA1 [105] and over 80 mutations identified in CNGB1 [106]. Most of these mutations are loss-of-function mutations. Mutations in CNGA3 and CNGB3 are associated with achromatopsia, progressive CRD, and early-onset macular degeneration [107,108,109]. There are over 100 and 40 mutations identified in CNGA3 and CNGB3, respectively, accounting for 70–80% of the total achromatopsia cases [108,110,111]. Mouse models with CNG channel deficiency mimic phenotypes in human patients, and they are manifested as reduced photoreceptor function and retinal degeneration [112,113,114]. Reduced [Ca2+]i and accumulation of cellular cGMP have been shown in the cones of CNG channel-deficient mice [113,114,115,116]. The cellular cGMP level increased at P8, peaked around P10–P15, remained high from P30–P60, and returned near the control level at P90 [113]. The abnormal accumulation of cGMP in photoreceptors is also demonstrated in mice with Cngb1 deficiency [117]. Unlike that in PDE6 deficiency, in which the impaired degradation of cGMP leads to its accumulation, CNG channel deficiency results in a reduction in [Ca2+]i, which subsequently leads to the overproduction of cGMP via the Ca2+/Mg2+-GCAP/RetGC complex.

3.4. Deficiency of Other Photoreceptor Specific Proteins

Interestingly, the accumulation of cGMP has been observed in rd2/rds mice with deficiencies of peripherin/Rds [49,118]. Peripherin/Rds is a disc membrane-integral protein. Dysfunction of this protein leads to the impaired morphogenesis of the outer segments, and it is associated with RP and macular degeneration [119,120]. How the dysfunction of peripherin/Rds leads to the accumulation of cGMP remains unknown currently. It might be related to the role of structural integrity of the outer segments in the biosynthesis/degradation of cGMP in photoreceptors.

4. Experimental Evidence Supporting the Contribution of Elevated cGMP Signaling to Photoreceptor Degeneration

Evidence supporting the contribution of the elevated cGMP signaling to photoreceptor degeneration is mainly obtained from the studies by depleting cGMP, deleting a CNG channel, or inhibiting PKG. All these approaches lead to photoreceptor protection in mouse models with the dysregulation of cGMP signaling.

4.1. Depletion of cGMP Reduces Photoreceptor Degeneration

The depletion of cellular cGMP reduces photoreceptor degeneration in mouse models with PDE6 deficiency and CNG channel deficiency. Knockdown of Gucy2e increases visual function and photoreceptor survival in Pde6b mutant mice [121]. The deletion of Gucy2e rescues the retinal phenotype in Cnga3−/− and Cnga3−/−/Nrl−/− mice (Cnga3-deficient mice on a cone dominant Nrl−/− background), manifested as increased cone density and with expression levels of cone-specific proteins [113,122,123]. The suppression of guanosine nucleotides/cGMP production, using mycophenolate mofetil (MMF), protects photoreceptors in rd10 and rd1 mice. MMF treatment significantly delays the onset of retinal degeneration, cGMP-dependent photoreceptor cytotoxicity, retinal/visual function, and microglial activation in rd10 and rd1 mice [124]. In addition, the overexpression of RD3 attenuates photoreceptor degeneration in transgenic mice expressing the human R838S RetGC1 dominant mutant [125]. The R838S RetGC1 mutation increases the enzyme’s affinity for Mg2+-GCAP (the activated form of GCAP) or makes it more difficult for RD3 to prevent the aberrant activation of RetGC by GCAP in the inner segment, leading to the increased synthesis of cGMP. The overexpression of RD3 leads to the suppression of RetGC1, as well as the rescue of the retinal phenotype observed in R838S RetGC1 mutant mice.

4.2. Inhibition of CNG Channel to Decrease Ca2+ Influx Reduces Photoreceptor Degeneration

The inhibition/deficiency of a CNG channel reduces retinal phenotypes in PDE6-deficient mice. Knockdown of Cnga1 increases visual function and photoreceptor survival in Pde6b mutant mice [121]. Similarly, the deletion of Cngb1 reduces photoreceptor death in Pde6g mutant mice [117]. Furthermore, the deletion of Cnga3 reduces cone degeneration in cpfl1 mice [126]. It should be pointed out that the cGMP level remains pathologically high in PDE6-deficient mice after the deletion of the CNG channel. However, photoreceptor viability and outer segment morphology are greatly improved, suggesting that deleterious Ca2+ influx is the main cause of the rapid death of photoreceptors [127].

4.3. Inhibition of PKG Reduces Photoreceptor Degeneration

The sole enzyme targets of cGMP are the PKGs. PKGs are serine/threonine kinases that are present in a variety of eukaryotes. There are two genes, Prkg1 and Prkg2, encoding for PKGs. The Prkg1 encodes for PKGI, whereas the Prkg2 encodes for PKGII [13,128,129]. The expression and activity of PKG are upregulated in the retinas of mice with abnormally elevated cellular cGMP levels [113,114,122]. In line with this, the inhibition of PKG reduces photoreceptor degeneration in mouse models with the dysregulation of cGMP signaling. Treatment with PKG inhibitors/cGMP analogues, including the classic compounds KT5823 and (RP)-8-Br-cGMPS, as well as the newly developed compounds CN03 and CN04, protects photoreceptors in PDE6-deficient mice, rd2/rds mice, and CNG channel-deficient mice [49,122,130]. The contribution of PKG to photoreceptor degeneration is also demonstrated by the genetic deletion of the enzymes. The deletion of Prkg2 reduces ER stress and cone death, resulting in long-lasting preservation for the morphology and structure of cone photoreceptors in Cnga3-deficient mice [131]. Additionally, the deletion of Prkg1 reduces rod degeneration in Cngb1-deficient mice [117].

5. The Cellular and Molecular Mechanisms Underlying cGMP Signaling-Induced Photoreceptor Degeneration

There are multiple mechanisms involved in cGMP dysregulation-induced photoreceptor degeneration (Figure 2). The elevated cGMP/PKG signaling impairs endoplasmic reticulum (ER) Ca2+ homeostasis and induces ER stress-associated cell death. Similarly, the low [Ca2+]i in CNG channel deficiency harms the ER Ca2+ homeostasis and deteriorates ER function. The elevated cGMP/PKG signaling also activates histone deacetylase (HDAC)/poly-ADP-ribose polymerase (PARP) signaling, leading to DNA condensation/damage and cell death. The elevated cGMP signaling and the subsequent cellular Ca2+ overload triggers the calpain-associated cell death pathway. Moreover, the ER stress, activation of PARP signaling, and activation of calpain signaling induce the mitochondrial insult and release of the apoptosis-inducing factor (AIF) to worsen the death process.

5.1. Elevated PKG Signaling

5.1.1. Excessive PKG Signaling Induces ER Stress

The ER is a large membrane-enclosed cellular organelle dedicated to biosynthesis, folding, and the assembly of membrane/secretory proteins, and it functions as a free Ca2+ storage. ER homeostasis is critical for its function, and impaired ER homeostasis will lead to unfolded protein response (UPR)/ER stress, impaired protein processing/trafficking, and impaired ER-associated protein degradation (ERAD). Under the stress conditions or accumulation of incorrectly folded/unfolded proteins, ER will undergo UPR/ER stress, which triggers cytotoxicity and cell death. ER stress is typically characterized by the activation of the three pathways/arms, including the pancreatic ER kinase (PKR)-like ER kinase (PERK)/eukaryotic-initiation factors 2α (phospho-eIF2α) pathway, the cleaved activating transcription factor 6 (ATF6) pathway, and the inositol-requiring kinase 1 (IRE1) pathway [132,133,134]. There are a variety of stress factors, originating from both cytosol and ER lumen, activating the three arms and initiating UPR, ER stress, and, ultimately, cell death.
Elevated PKG signaling impairs ER homeostasis and induces ER stress-associated photoreceptor death (Figure 2). This cellular event is well characterized in mice with CNG channel deficiency. Mice with cone CNG channel deficiency show ER stress-associated cone death, manifested as increased ER stress marker proteins, including Grp78/Bip, phospho-eIF2, phospho-IP3R, and CCAAT/enhancer-binding protein homologous protein (CHOP), along with impaired protein trafficking and impaired ERAD [114,122,123,135]. ER stress-associated cell death in CNG channel deficiency is also manifested by the up-regulation of the cysteine protease calpains and the cleavage of the cysteine proteases, caspase-12 and caspase-7. Caspase-12 is predominantly localized at the ER, activated upon ER stress and Ca2+ release from the ER store, and subsequently translocated into the nucleus to induce apoptosis [136,137]. Caspase-7 resides in the ER neighborhood along with caspase-12 [138]. It is activated under ER stress, and the activated caspase-7 cleaves procaspase-12 to generate active forms of the protease [138]. The involvement of caspase-12, and its translocation to the nucleus, is also demonstrated in Pde6b mutant retinas [139]. Treatment with a chemical chaperone (tauroursodeoxycholic acid, TUDCA) or a molecular chaperone (11-cis-retinal) reduces ER stress responses/cone death in CNG channel deficiency and improves protein trafficking/outer segment location [123,135], demonstrating a role of ER stress in the photoreceptor degeneration. The link between activated PKG signaling and ER stress/cell death in CNG channel deficiency is demonstrated by the depletion of cGMP/the deletion of RetGC1 and inhibition of PKG using PKG inhibitors and PKG deletions. The depletion of cGMP or inhibition of PKG significantly reduces ER stress/cone death and improves protein trafficking to the outer segments in CNG channel-deficient mice [113,123,135].

5.1.2. Excessive PKG Signaling Impairs ER Ca2+ Homeostasis

As a storage of free Ca2+, ER plays a pivotal role in cellular/ER Ca2+ homeostasis. ER Ca2+ homeostasis and regulation are vital for ER to function properly, including protein folding/trafficking. ER Ca2+ homeostasis is governed by the ER Ca2+ channels, inositol 1,4,5-trisphosphate receptors (IP3Rs) and ryanodine receptors (RyRs), on the ER membrane, for efflux out of the ER into cytosol, and the sarco/endoplasmic reticulum Ca2+-ATPase (SERCA) for influx into the ER. There are three IP3R isoforms: IP3R1, IP3R2, and IP3R3. In the mouse retina, IP3R1 is the major IP3R isoform, its mRNA level is approximately 6 to 10-fold higher than the mRNA level of IP3R3, whereas IP3R2 mRNA is not detectable in transcript expression studies [123]. There are three RyR isoforms: RyR1, RyR2, and RyR3 [140]. All three isoforms of RyRs are expressed in photoreceptors, and RyR2 is identified as the major form in the photoreceptors [140]. The activities of ER Ca2+ channels are highly regulated by their respective ligands, IP3 and ryanodine, cytosolic Ca2+ levels, and other regulatory signaling, including PKG signaling.
The impairment of ER Ca2+ homeostasis in the dysregulation of PKG signaling is well characterized in mice with CNG channel deficiency (Figure 2). The expression/activity of IP3R1 and RyR2 is significantly increased in these mice, along with protein mistrafficking/mislocalization, ER stress, and impaired ERAD [123]. The inhibition of PKG by chemical inhibitors or genetic deletion, or the depletion of cGMP by Gucy2e deletion, reduces the expression/activity of IP3R1 and RyR2, along with reduced ER stress/cone death and improved protein trafficking/outer segment localization [122,131,141,142]. The regulation of PKG on the ER calcium channels is also shown in a cell culture model system. Treatment with the cGMP analogue significantly enhances mRNA levels of IP3R1 and RYR2 in cultured photoreceptor-derived Weri-Rb1 cells [141]. Furthermore, the inhibition of IP3R1 and RyR2, by chemical inhibitors or genetic deletion, reduces ER stress/photoreceptor death and improves protein trafficking/outer segment localization and ERAD [123,142,143]. Thus, the elevated PKG signaling stimulates ER Ca2+ channels, leading to increased ER Ca2+ release/depletion and impaired ER function/processing.

5.1.3. Other Potential Targets of PKG in Photoreceptor Degeneration

As a serine–threonine kinase, PKGs have numerous phosphorylation protein targets. Although the regulation of PKG signaling in ER stress/ER Ca2+ homeostasis has been characterized in CNG channel deficiency, the substrates of PKG and the downstream pathways triggered by its activation in retinal degeneration remain not fully understood. A recent kinome activity-profiling study, which applies multiplex peptide microarrays to identify proteins whose phosphorylation are significantly altered by PKG activation, has identified several potential PKG target proteins in rd1 mouse retinal explants [144]. Among the PKG substrates are potassium channels belonging to the Kv1 family (KCNA3, KCNA6), cyclic AMP-responsive element-binding protein 1 (CREB1), DNA topoisomerase 2-α (TOP2A), 6-phosphofructo-2-kinase/fructose-2,6-biphosphatase 3 (F263), and the glutamate ionotropic receptor kainate 2 (GRIK2). The retinal expression of these PKG targets is confirmed by immunofluorescence labeling [144]. The elevated CREB signaling is also observed in CNG channel-deficient retina, and the deletion of Gucy2e reverses this alteration [141]. In addition, the transcriptome analysis of rd1 retinal explants suggests that PKG signaling negatively affects oxidative phosphorylation and mitochondrial function [145]. However, whether these molecules contribute to photoreceptor degeneration remains to be investigated.

5.2. Impaired Cellular Ca2+ Homeostasis

5.2.1. Ca2+ Overload and Calpain Activation

In PDE6 deficiency or overactivation of RetGC1, the accumulation of cGMP will lead to a constitutive opening of the CNG channel and intracellular Ca2+ overload (Figure 2). The most prominent consequence of Ca2+ overload is the activation of the calcium-activated cysteine protease calpains. Calpains exist as inactive proenzymes in the cytosol. When the intracellular Ca2+ level is high, it triggers the conversion of the proenzymes to their active forms. Activated calpains then cleave a variety of cytoplasmic and nuclear substrates, including caspase 3 and PARP [146,147]. The activation of calpains also induces the release of AIF from mitochondria, leading to prolonged proteolysis and apoptosis/necroptosis [114,139,148,149,150,151,152]. Retinas of rd1 and rd10 mice show increased calpain activity, which peaks synchronously with cell death [139,148,150,152]. An elevation of calpain activity is also observed in Cpfl1 mice [98]. The contribution of calpains to cell death in retinal degeneration is demonstrated by using the calpain-specific inhibitors. Treatment with calpain inhibitor XI significantly reduces cell death in rd1 and rd10 mice [148,149]. Interestingly, the contribution of calpain activity to photoreceptor death is also shown in an induced animal model of retinal degeneration. The activation of calpains and the calpain-specific proteolysis of α-spectrin are observed in experimental rats treated with N-methyl-N-nitrosourea (MNU), an alkylating agent that interacts with DNA, specifically targets retinal photoreceptors, and causes oxidative stress/damage. The administration of a calpain inhibitor induces a significant protective effect against photoreceptor loss in MNU-treated rats [153].

5.2.2. Ca2+ Depletion

Ca2+ homeostasis is crucial for normal cell function. Ca2+ overload is cytotoxic. Likewise, Ca2+ depletion also renders harmful effects. Reduced [Ca2+]i has been shown in the cones of CNG channel-deficient mice [123]. Similar pathophysiological conditions may be present in LCA-RetGC mutation/RetGC1 deficiency [19,61,154], as well as in a rd3 mutant mouse [155]. When the cytosolic Ca2+ level is low/reduced, it will affect not only the function of many Ca2+ binding/regulating proteins but also the function/homeostasis of ER and mitochondria. More profoundly, when the CNG channel is intact, the reduced cellular Ca2+ level will activate GCAP/RetGC, leading to the accumulation of cGMP and subsequent activation of PKG (Figure 2).

5.3. Other Factors

5.3.1. HDAC and PARP

HDACs are enzymes that catalyze the removal of acetyl functional groups from the lysine residues of both histone and nonhistone proteins. Excessive activation of HDAC is associated with photoreceptor death in mouse models with cGMP signaling perturbation (Figure 2). HDAC activity is increased in the retinas of rd1 and Cpfl1 mice, and treatment with HDAC inhibitors reduces photoreceptor degeneration in these mice [156,157]. Nevertheless, how HDAC is activated in the retinas of these mice remains less understood. Based on the comparative localization studies [158], PKG might be the upstream activator. PARP is a DNA-repairing enzyme that is involved in genomic stability and programmed cell death through the formation of poly (ADP-ribose) polymers (PAR), and its activation causes a parthanatos-related form of cell death [159]. Rd1 photoreceptors actively undergoing cell death show a great elevation of PARP activity. The elevated PARP is co-labeled with oxidatively damaged DNA and the nuclear translocation of AIF. Moreover, treatment with the PARP inhibitor reduces cell death in rd1 retinal explants [160]. The association of PARP activation with HDAC is demonstrated by using HDAC inhibitors. Treatment with HDAC inhibitors reduces PARP activation and photoreceptor degeneration in rd1 mice [156]. In addition to HDAC, PARP activation may result from elevated cellular Ca2+ and the activation of calpain [161,162]. PARP is cleaved, by purified calpain, into fragments, suggesting it might be one of the calpain substrates [147].

5.3.2. AIF

AIF is the protein that triggers chromatin condensation and DNA degradation. It resides, normally, within the intermembrane space of mitochondria. Upon certain death stimuli/mitochondrial damage, AIF translocates into the cytosol and, ultimately, the nucleus, where it contributes to DNA fragmentation and chromatin condensation, initiating a caspase-independent cell death process [163,164]. Nuclear translocation of AIF is observed in the retinas of CNG channel-deficient mice [114] and rd1 mice [160], suggesting a potential role of mitochondrial insult and nuclear translocation of AIF in retinal degeneration with the dysregulation of cGMP signaling. The activation of calpain and PARP has been shown to be associated with the activation of AIF [165,166,167]. In rd1 mice, the nuclear translocation of AIF is overlapped with PARP activity staining [160], suggesting that PARP mediates the translocation of AIF. Furthermore, calpain inhibitors interfere with the AIF activation/nuclear translocation in the rd1 retina [139], supporting an involvement of calpain in the activation of AIF.

6. Summary and Perspectives

Photoreceptor cGMP level is tightly controlled by a well-developed, feedback regulatory system. The Ca2+/Mg2+-GCAP/RetGC complex, with the modulation of RD3, governs the biogeneration of cGMP, whereas PDE6 is the sole force for degradation of cGMP. The binding of cGMP to the CNG channel regulates cellular Ca2+ levels, membrane potential, and phototransduction. As second messengers, cGMP and Ca2+ regulate a variety of cellular activities via their downstream targets. The dysregulation of cellular cGMP homeostasis leads to impaired phototransduction and photoreceptor degeneration, and it is associated with retinal diseases, including RP, LCA, CRD, and achromatopsia. Overactivation of RetGC/GCAP, deficiency of RD3, or deficiency of PDE6 lead to an accumulation of cGMP. The deficiency of a CNG channel leads to the accumulation of cGMP via the Ca2+/Mg2+-GCAP/RetGC complex. The deficiency of a CNG channel reduces Ca2+ influx, leading to reduced [Ca2+]i and the subsequent activation of RetGC. Activated cGMP/PKG signaling induces ER stress and the impairment of ER Ca2+ homeostasis. Either cellular Ca2+ overload or Ca2+ depletion will impair cellular homeostasis, leading to cell stress/death. In PDE6 deficiency, RD3 deficiency, or RetGC overaction, the activated cGMP/PKG signaling, along with Ca2+ overload/calpain activation, plays a significant role in the disease pathogenesis and cell death progression. In CNG channel deficiency or RetGC deficiency, the activated cGMP/PKG signaling, along with cellular Ca2+ depletion, plays a significant role in the disease pathogenesis and cell death progression. The activated cGMP/PKG signaling may also induce HDAC and PARP, leading to cell stress/death directly and via mitochondrial dysfunction/the release of AIF. Thus, targeting cGMP/PKG signaling in photoreceptors and the downstream target cellular organelles/molecules represents a significant approach to slow photoreceptor death in retinal degenerative diseases.

Funding

This work was supported by grants from the National Eye Institute (R01EY027754, R01EY033841, and P30EY021725) and the Oklahoma Center for the Advancement of Science and Technology.

Conflicts of Interest

All authors declare that they have no conflict of interest.

References

  1. Kawamura, S.; Tachibanaki, S. Rod and cone photoreceptors: Molecular basis of the difference in their physiology. Comp. Biochem. Physiol. Part A Mol. Integr. Physiol. 2008, 150, 369–377. [Google Scholar] [CrossRef] [PubMed]
  2. Korenbrot, J.I. Speed, sensitivity, and stability of the light response in rod and cone photoreceptors: Facts and models. Prog. Retin. Eye Res. 2012, 31, 442–466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Burns, M.E.; Arshavsky, V.Y. Beyond Counting Photons: Trials and Trends in Vertebrate Visual Transduction. Neuron 2005, 48, 387–401. [Google Scholar] [CrossRef] [Green Version]
  4. Ebrey, T.; Koutalos, Y. Vertebrate photoreceptors. Prog. Retin. Eye Res. 2001, 20, 49–94. [Google Scholar] [CrossRef] [PubMed]
  5. Molday, R.S.; Moritz, O.L. Photoreceptors at a glance. J. Cell Sci. 2015, 128, 4039–4045. [Google Scholar] [CrossRef] [Green Version]
  6. Barret, D.C.A.; Kaupp, U.B.; Marino, J. The structure of cyclic nucleotide-gated channels in rod and cone photoreceptors. Trends Neurosci. 2022, 45, 763–776. [Google Scholar] [CrossRef]
  7. Leskov, I.B.; Klenchin, V.A.; Handy, J.W.; Whitlock, G.G.; Govardovskii, V.I.; Bownds, M.D.; Lamb, T.D.; Pugh, E.N., Jr.; Arshavsky, V.Y. The gain of rod phototransduction: Reconciliation of biochemical and electrophysiological measurements. Neuron 2000, 27, 525–537. [Google Scholar] [CrossRef] [Green Version]
  8. Arshavsky, V.Y.; Lamb, T.D.; Pugh, E.N., Jr. G Proteins and Phototransduction. Annu. Rev. Physiol. 2002, 64, 153–187. [Google Scholar] [CrossRef] [Green Version]
  9. Stryer, L. Cyclic GMP cascade of vision. Annu. Rev. Neurosci. 1986, 9, 87–119. [Google Scholar] [CrossRef]
  10. Kaupp, U.B.; Seifert, R. Cyclic nucleotide-gated ion channels. Physiol. Rev. 2002, 82, 769–824. [Google Scholar] [CrossRef] [Green Version]
  11. Deng, W.T.; Kolandaivelu, S.; Dinculescu, A.; Li, J.; Zhu, P.; Chiodo, V.A.; Ramamurthy, V.; Hauswirth, W.W. Cone Phosphodiesterase-6gamma’ Subunit Augments Cone PDE6 Holoenzyme Assembly and Stability in a Mouse Model Lacking Both Rod and Cone PDE6 Catalytic Subunits. Front. Mol. Neurosci. 2018, 11, 233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Lagman, D.; Franzen, I.E.; Eggert, J.; Larhammar, D.; Abalo, X.M. Evolution and expression of the phosphodiesterase 6 genes unveils vertebrate novelty to control photosensitivity. BMC Evol. Biol. 2016, 16, 124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Tolone, A.; Belhadj, S.; Rentsch, A.; Schwede, F.; Paquet-Durand, F. The cGMP Pathway and Inherited Photoreceptor Degeneration: Targets, Compounds, and Biomarkers. Genes 2019, 10, 453. [Google Scholar] [CrossRef] [Green Version]
  14. Paquet-Durand, F.; Marigo, V.; Ekstrom, P. RD Genes Associated with High Photoreceptor cGMP-Levels (Mini-Review). Adv. Exp. Med. Biol. 2019, 1185, 245–249. [Google Scholar]
  15. Francis, S.H.; Blount, M.A.; Corbin, J.D. Mammalian Cyclic Nucleotide Phosphodiesterases: Molecular Mechanisms and Physiological Functions. Physiol. Rev. 2011, 91, 651–690. [Google Scholar] [CrossRef] [Green Version]
  16. Dizhoor, A.M.; Lowe, D.G.; Olshevskaya, E.V.; Laura, R.P.; Hurley, J.B. The human photoreceptor membrane guanylyl cyclase, RetGC, is present in outer segments and is regulated by calcium and a soluble activator. Neuron 1994, 12, 1345–1352. [Google Scholar] [CrossRef] [PubMed]
  17. Lim, S.; Roseman, G.; Peshenko, I.; Manchala, G.; Cudia, D.; Dizhoor, A.; Millhauser, G.; Ames, J.B. Retinal guanylyl cyclase activating protein 1 forms a functional dimer. PLoS ONE 2018, 13, e0193947. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Yang, R.-B.; Fülle, H.-J.; Garbers, D.L. Chromosomal Localization and Genomic Organization of Genes Encoding Guanylyl Cyclase Receptors Expressed in Olfactory Sensory Neurons and Retina. Genomics 1996, 31, 367–372. [Google Scholar] [CrossRef]
  19. Baehr, W.; Karan, S.; Maeda, T.; Luo, D.-G.; Li, S.; Bronson, J.D.; Watt, C.B.; Yau, K.-W.; Frederick, J.M.; Palczewski, K. The Function of Guanylate Cyclase 1 and Guanylate Cyclase 2 in Rod and Cone Photoreceptors. J. Biol. Chem. 2007, 282, 8837–8847. [Google Scholar] [CrossRef] [Green Version]
  20. Peshenko, I.V.; Olshevskaya, E.V.; Savchenko, A.B.; Karan, S.; Palczewski, K.; Baehr, W.; Dizhoor, A.M. Enzy-matic properties and regulation of the native isozymes of retinal membrane guanylyl cyclase (RetGC) from mouse photoreceptors. Biochemistry 2011, 50, 5590–5600. [Google Scholar] [CrossRef] [Green Version]
  21. Lowe, D.G.; Dizhoor, A.M.; Liu, K.; Gu, Q.; Spencer, M.; Laura, R.; Lu, L.; Hurley, J.B. Cloning and expression of a second photoreceptor-specific membrane retina guanylyl cyclase (RetGC), RetGC-2. Proc. Natl. Acad. Sci. USA 1995, 92, 5535–5539. [Google Scholar] [CrossRef] [PubMed]
  22. Peshenko, I.V.; Olshevskaya, E.V.; Dizhoor, A.M. Dimerization Domain of Retinal Membrane Guanylyl Cyclase 1 (RetGC1) Is an Essential Part of Guanylyl Cyclase-activating Protein (GCAP) Binding Interface. J. Biol. Chem. 2015, 290, 19584–19596. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Peshenko, I.V.; Olshevskaya, E.V.; Dizhoor, A.M. Evaluating the Role of Retinal Membrane Guanylyl Cyclase 1 (RetGC1) Domains in Binding Guanylyl Cyclase-activating Proteins (GCAPs). J. Biol. Chem. 2015, 290, 6913–6924. [Google Scholar] [CrossRef] [Green Version]
  24. Dizhoor, A.M.; Peshenko, I.V. Regulation of retinal membrane guanylyl cyclase (RetGC) by negative calcium feedback and RD3 protein. Pflug. Arch. 2021, 473, 1393–1410. [Google Scholar] [CrossRef]
  25. Liu, Y.; Ruoho, A.E.; Rao, V.D.; Hurley, J.H. Catalytic mechanism of the adenylyl and guanylyl cyclases: Modeling and mutational analysis. Proc. Natl. Acad. Sci. USA 1997, 94, 13414–13419. [Google Scholar] [CrossRef] [PubMed]
  26. Olshevskaya, E.V.; Peshenko, I.V.; Savchenko, A.B.; Dizhoor, A.M. Retinal Guanylyl Cyclase Isozyme 1 Is the Preferential In Vivo Target for Constitutively Active GCAP1 Mutants Causing Congenital Degeneration of Photoreceptors. J. Neurosci. 2012, 32, 7208–7217. [Google Scholar] [CrossRef] [Green Version]
  27. Makino, C.L.; Peshenko, I.V.; Wen, X.H.; Olshevskaya, E.V.; Barrett, R.; Dizhoor, A.M. A role for GCAP2 in regulating the photoresponse. Guanylyl cyclase activation and rod electrophysiology in GUCA1B knock-out mice. J. Biol. Chem. 2008, 283, 29135–29143. [Google Scholar] [CrossRef] [Green Version]
  28. López-Begines, S.; Plana-Bonamaisó, A.; Méndez, A. Molecular determinants of Guanylate Cyclase Activating Protein subcellular distribution in photoreceptor cells of the retina. Sci. Rep. 2018, 8, 2903. [Google Scholar] [CrossRef] [Green Version]
  29. Dizhoor, A. Regulation of cGMP synthesis in photoreceptors: Role in signal transduction and congenital diseases of the retina. Cell. Signal. 2000, 12, 711–719. [Google Scholar] [CrossRef]
  30. Ames, J.B.; Dizhoor, A.M.; Ikura, M.; Palczewski, K.; Stryer, L. Three-dimensional structure of guanylyl cyclase activating protein-2, a calcium-sensitive modulator of photoreceptor guanylyl cyclases. J. Biol. Chem. 1999, 274, 19329–19337. [Google Scholar] [CrossRef] [Green Version]
  31. Makino, C.L.; Wen, X.-H.; Olshevskaya, E.V.; Peshenko, I.V.; Savchenko, A.B.; Dizhoor, A.M. Enzymatic Relay Mechanism Stimulates Cyclic GMP Synthesis in Rod Photoresponse: Biochemical and Physiological Study in Guanylyl Cyclase Activating Protein 1 Knockout Mice. PLoS ONE 2012, 7, e47637. [Google Scholar] [CrossRef]
  32. Ames, J.B. Structural Insights into Retinal Guanylate Cyclase Activator Proteins (GCAPs). Int. J. Mol. Sci. 2021, 22, 8731. [Google Scholar] [CrossRef] [PubMed]
  33. Peshenko, I.V.; Yu, Q.; Lim, S.; Cudia, D.; Dizhoor, A.M.; Ames, J.B. Retinal degeneration 3 (RD3) protein, a retinal guanylyl cyclase regulator, forms a monomeric and elongated four-helix bundle. J. Biol. Chem. 2019, 294, 2318–2328. [Google Scholar] [CrossRef] [Green Version]
  34. Dizhoor, A.M.; Olshevskaya, E.V.; Peshenko, I.V. Retinal degeneration-3 protein promotes photoreceptor survival by suppressing activation of guanylyl cyclase rather than accelerating GMP recycling. J. Biol. Chem. 2021, 296, 100362. [Google Scholar] [CrossRef] [PubMed]
  35. Azadi, S.; Molday, L.L.; Molday, R.S. RD3, the protein associated with Leber congenital amaurosis type 12, is required for guanylate cyclase trafficking in photoreceptor cells. Proc. Natl. Acad. Sci. USA 2010, 107, 21158–21163. [Google Scholar] [CrossRef]
  36. Conti, M.; Beavo, J. Biochemistry and Physiology of Cyclic Nucleotide Phosphodiesterases: Essential Components in Cyclic Nucleotide Signaling. Annu. Rev. Biochem. 2007, 76, 481–511. [Google Scholar] [CrossRef]
  37. Cote, R.H.; Gupta, R.; Irwin, M.J.; Wang, X. Photoreceptor Phosphodiesterase (PDE6): Structure, Regulatory Mechanisms, and Implications for Treatment of Retinal Diseases. Adv. Exp. Med. Biol. 2022, 1371, 33–59. [Google Scholar]
  38. Cote, R.H. Characteristics of Photoreceptor PDE (PDE6): Similarities and differences to PDE5. Int. J. Impot. Res. 2004, 16 (Suppl. S1), S28–S33. [Google Scholar] [CrossRef] [Green Version]
  39. Cote, R.H. Photoreceptor phosphodiesterase (PDE6): Activation and inactivation mechanisms during visual transduction in rods and cones. Pflug. Arch. 2021, 473, 1377–1391. [Google Scholar] [CrossRef] [PubMed]
  40. Barren, B.; Gakhar, L.; Muradov, H.; Boyd, K.K.; Ramaswamy, S.; Artemyev, N.O. Structural basis of phosphodiesterase 6 inhibition by the C-terminal region of the gamma-subunit. Embo J. 2009, 28, 3613–3622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Biel, M.; Michalakis, S. Cyclic nucleotide-gated channels. In Handbook of Experimental Pharmacology; Springer: Berlin/Heidelberg, Germany, 2009; pp. 111–136. [Google Scholar]
  42. Zheng, J.; Trudeau, M.C.; Zagotta, W.N. Rod Cyclic Nucleotide-Gated Channels Have a Stoichiometry of Three CNGA1 Subunits and One CNGB1 Subunit. Neuron 2002, 36, 891–896. [Google Scholar] [CrossRef] [Green Version]
  43. Weitz, D.; Ficek, N.; Kremmer, E.; Bauer, P.J.; Kaupp, U. Subunit Stoichiometry of the CNG Channel of Rod Photoreceptors. Neuron 2002, 36, 881–889. [Google Scholar] [CrossRef] [Green Version]
  44. Zheng, X.; Hu, Z.; Li, H.; Yang, J. Structure of the human cone photoreceptor cyclic nucleotide-gated channel. Nat. Struct. Mol. Biol. 2022, 29, 40–46. [Google Scholar] [CrossRef]
  45. Ding, X.Q.; Matveev, A.; Singh, A.; Komori, N.; Matsumoto, H. Biochemical characterization of cone cyclic nucle-otide-gated (CNG) channel using the infrared fluorescence detection system. Adv. Exp. Med. Biol. 2012, 723, 769–775. [Google Scholar] [PubMed] [Green Version]
  46. Kesters, D.; Brams, M.; Nys, M.; Wijckmans, E.; Spurny, R.; Voets, T.; Tytgat, J.; Kusch, J.; Ulens, C. Structure of the SthK Carboxy-Terminal Region Reveals a Gating Mechanism for Cyclic Nucleotide-Modulated Ion Channels. PLoS ONE 2015, 10, e0116369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Korenbrot, J.I.; Rebrik, T.I. Tuning Outer Segment Ca2+Homeostasis to Phototransduction in Rods and Cones. Adv. Exp. Med. Biol. 2002, 514, 179–203. [Google Scholar] [CrossRef]
  48. Michalakis, S.; Becirovic, E.; Biel, M. Retinal Cyclic Nucleotide-Gated Channels: From Pathophysiology to Therapy. Int. J. Mol. Sci. 2018, 19, 749. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Paquet-Durand, F.; Hauck, S.M.; van Veen, T.; Ueffing, M.; Ekström, P. PKG activity causes photoreceptor cell death in two retinitis pigmentosa models. J. Neurochem. 2009, 108, 796–810. [Google Scholar] [CrossRef]
  50. Kelsell, R.E.; Gregory-Evans, K.; Payne, A.; Perrault, I.; Kaplan, J.; Yang, R.-B.; Garbers, D.L.; Bird, A.C.; Moore, A.T.; Hunt, D.M. Mutations in the Retinal Guanylate Cyclase (RETGC-1) Gene in Dominant Cone-Rod Dystrophy. Hum. Mol. Genet. 1998, 7, 1179–1184. [Google Scholar] [CrossRef]
  51. Tucker, C.L.; Woodcock, S.C.; Kelsell, R.E.; Ramamurthy, V.; Hunt, D.M.; Hurley, J.B. Biochemical analysis of a dimerization domain mutation in RetGC-1 associated with dominant cone–rod dystrophy. Proc. Natl. Acad. Sci. USA 1999, 96, 9039–9044. [Google Scholar] [CrossRef]
  52. Duda, T.; Koch, K.W. Retinal diseases linked with photoreceptor guanylate cyclase. Mol. Cell Biochem. 2002, 230, 129–138. [Google Scholar] [CrossRef] [PubMed]
  53. Kitiratschky, V.B.; Wilke, R.; Renner, A.B.; Kellner, U.; Vadala, M.; Birch, D.G.; Wissinger, B.; Zrenner, E.; Kohl, S. Mutation analysis identifies GUCY2D as the major gene responsible for autosomal dominant progressive cone degeneration. Investig. Ophthalmol. Vis. Sci. 2008, 49, 5015–5023. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Sharon, D.; Wimberg, H.; Kinarty, Y.; Koch, K.-W. Genotype-functional-phenotype correlations in photoreceptor guanylate cyclase (GC-E) encoded by GUCY2D. Prog. Retin. Eye Res. 2018, 63, 69–91. [Google Scholar] [CrossRef] [PubMed]
  55. Wimberg, H.; Lev, D.; Yosovich, K.; Namburi, P.; Banin, E.; Sharon, D.; Koch, K.-W. Photoreceptor Guanylate Cyclase (GUCY2D) Mutations Cause Retinal Dystrophies by Severe Malfunction of Ca2+-Dependent Cyclic GMP Synthesis. Front. Mol. Neurosci. 2018, 11, 348. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Rozet, J.-M.; Perrault, I.; Gerber, S.; Hanein, S.; Barbet, F.; Ducroq, D.; Souied, E.; Munnich, A.; Kaplan, J. Complete abolition of the retinal-specific guanylyl cyclase (retGC-1) catalytic ability consistently leads to leber congenital amaurosis (LCA). Investig. Ophthalmol. Vis. Sci. 2001, 42, 1190–1192. [Google Scholar]
  57. Tucker, C.; Ramamurthy, V.; Pina, A.-L.; Loyer, M.; Dharmaraj, S.; Li, Y.; Maumenee, I.H.; Hurley, J.B.; Koenekoop, R.K. Functional analyses of mutant recessive GUCY2D alleles identified in Leber congenital amaurosis patients: Protein domain comparisons and dominant negative effects. Mol. Vis. 2004, 10, 297–303. [Google Scholar]
  58. Yi, Z.; Sun, W.; Xiao, X.; Li, S.; Jia, X.; Li, X.; Yu, B.; Wang, P.; Zhang, Q. Novel variants in GUCY2D causing retinopathy and the genotype-phenotype correlation. Exp. Eye Res. 2021, 208, 108637. [Google Scholar] [CrossRef]
  59. Liu, X.; Fujinami, K.; Kuniyoshi, K.; Kondo, M.; Ueno, S.; Hayashi, T.; Mochizuki, K.; Kameya, S.; Yang, L.; Fujinami-Yokokawa, Y.; et al. Clinical and Genetic Characteristics of 15 Affected Patients From 12 Japanese Families with GUCY2D-Associated Retinal Disorder. Transl. Vis. Sci. Technol. 2020, 9, 2. [Google Scholar] [CrossRef]
  60. Boye, S.E. Leber Congenital Amaurosis Caused by Mutations in GUCY2D. Cold Spring Harb. Perspect. Med. 2014, 5, a017350. [Google Scholar] [CrossRef] [Green Version]
  61. Jacobson, S.G.; Cideciyan, A.V.; Peshenko, I.V.; Sumaroka, A.; Olshevskaya, E.V.; Cao, L.; Schwartz, S.B.; Roman, A.J.; Olivares, M.B.; Sadigh, S.; et al. Determining consequences of retinal membrane guanylyl cyclase (RetGC1) deficiency in human Leber congenital amaurosis en route to therapy: Residual cone-photoreceptor vision correlates with biochemical properties of the mutants. Hum. Mol. Genet. 2013, 22, 168–183. [Google Scholar] [CrossRef] [Green Version]
  62. Dizhoor, A.M.; Olshevskaya, E.V.; Peshenko, I.V. The R838S Mutation in Retinal Guanylyl Cyclase 1 (RetGC1) Alters Calcium Sensitivity of cGMP Synthesis in the Retina and Causes Blindness in Transgenic Mice. J. Biol. Chem. 2016, 291, 24504–24516. [Google Scholar] [CrossRef] [Green Version]
  63. Sato, S.; Peshenko, I.V.; Olshevskaya, E.V.; Kefalov, V.J.; Dizhoor, A.M. GUCY2D Cone-Rod Dystrophy-6 Is a “Phototransduction Disease” Triggered by Abnormal Calcium Feedback on Retinal Membrane Guanylyl Cyclase 1. J. Neurosci. 2018, 38, 2990–3000. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Peshenko, I.V.; Olshevskaya, E.V.; Dizhoor, A.M. GUCY2D mutations in retinal guanylyl cyclase 1 provide bio-chemical reasons for dominant cone-rod dystrophy but not for stationary night blindness. J. Biol. Chem. 2020, 295, 18301–18315. [Google Scholar] [CrossRef] [PubMed]
  65. Jacobson, S.G.; Cideciyan, A.V.; Ho, A.C.; Roman, A.J.; Wu, V.; Garafalo, A.V.; Sumaroka, A.; Krishnan, A.K.; Swider, M.; Mascio, A.A.; et al. Night vision restored in days after decades of congenital blindness. iScience 2022, 25, 105274. [Google Scholar] [CrossRef] [PubMed]
  66. Jacobson, S.G.; Cideciyan, A.V.; Ho, A.C.; Peshenko, I.V.; Garafalo, A.V.; Roman, A.J.; Sumaroka, A.; Wu, V.; Krishnan, A.K.; Sheplock, R.; et al. Safety and improved efficacy signals following gene therapy in childhood blindness caused by GUCY2D mutations. iScience 2021, 24, 102409. [Google Scholar] [CrossRef]
  67. Ugur Iseri, S.A.; Durlu, Y.K.; Tolun, A. A novel recessive GUCY2D mutation causing cone-rod dystrophy and not Leber’s congenital amaurosis. Eur. J. Hum. Genet. 2010, 18, 1121–1126. [Google Scholar] [CrossRef]
  68. Stunkel, M.L.; Brodie, S.; Cideciyan, A.V.; Pfeifer, W.L.; Kennedy, E.L.; Stone, E.M.; Jacobson, S.G.; Drack, A.V. Expanded Retinal Disease Spectrum Associated with Autosomal Recessive Mutations in GUCY2D. Am. J. Ophthalmol. 2018, 190, 58–68. [Google Scholar] [CrossRef]
  69. Friedman, J.S.; Chang, B.; Kannabiran, C.; Chakarova, C.; Singh, H.P.; Jalali, S.; Hawes, N.L.; Branham, K.; Othman, M.; Filippova, E.; et al. Premature Truncation of a Novel Protein, RD3, Exhibiting Subnuclear Localization Is Associated with Retinal Degeneration. Am. J. Hum. Genet. 2006, 79, 1059–1070. [Google Scholar] [CrossRef] [Green Version]
  70. Chang, B.; Heckenlively, J.R.; Hawes, N.L.; Roderick, T.H. New Mouse Primary Retinal Degeneration (rd-3). Genomics 1993, 16, 45–49. [Google Scholar] [CrossRef]
  71. Dell’orco, D.; Cortivo, G.D. Normal GCAPs partly compensate for altered cGMP signaling in retinal dystrophies associated with mutations in GUCA1A. Sci. Rep. 2019, 9, 20105. [Google Scholar] [CrossRef] [Green Version]
  72. Biasi, A.; Marino, V.; Cortivo, G.D.; Maltese, P.E.; Modarelli, A.M.; Bertelli, M.; Colombo, L.; Dell’orco, D. A Novel GUCA1A Variant Associated with Cone Dystrophy Alters cGMP Signaling in Photoreceptors by Strongly Interacting with and Hyperactivating Retinal Guanylate Cyclase. Int. J. Mol. Sci. 2021, 22, 10809. [Google Scholar] [CrossRef]
  73. Manes, G.; Mamouni, S.; Hérald, E.; Richard, A.-C.; Sénéchal, A.; Aouad, K.; Bocquet, B.; Meunier, I.; Hamel, C.P. Cone dystrophy or macular dystrophy associated with novel autosomal dominant GUCA1A mutations. Mol. Vis. 2017, 23, 198–209. [Google Scholar]
  74. Vocke, F.; Weisschuh, N.; Marino, V.; Malfatti, S.; Jacobson, S.G.; Reiff, C.M.; Dell’Orco, D.; Koch, K.W. Dys-function of cGMP signalling in photoreceptors by a macular dystrophy-related mutation in the calcium sensor GCAP1. Hum. Mol. Genet. 2017, 26, 133–144. [Google Scholar]
  75. Michaelides, M.; Wilkie, S.E.; Jenkins, S.; Holder, G.E.; Hunt, D.M.; Moore, A.T.; Webster, A.R. Mutation in the Gene GUCA1A, Encoding Guanylate Cyclase-Activating Protein 1, Causes Cone, Cone-Rod, and Macular Dystrophy. Ophthalmology 2005, 112, 1442–1447. [Google Scholar] [CrossRef]
  76. Payne, A.; Downes, S.M.; Bessant, D.A.; Taylor, R.; Holder, G.E.; Warren, M.; Bird, A.C.; Bhattacharya, S.S. A mutation in guanylate cyclase activator 1A (GUCA1A) in an autosomal dominant cone dystrophy pedigree mapping to a new locus on chromosome 6p21.1. Hum. Mol. Genet. 1998, 7, 273–277. [Google Scholar] [CrossRef]
  77. Jiang, L.; Katz, B.J.; Yang, Z.; Zhao, Y.; Faulkner, N.; Hu, J.; Baird, J.; Baehr, W.; Creel, D.J.; Zhang, K. Autosomal dominant cone dystrophy caused by a novel mutation in the GCAP1 gene (GUCA1A). Mol. Vis. 2005, 11, 143–151. [Google Scholar]
  78. Olshevskaya, E.V.; Calvert, P.D.; Woodruff, M.L.; Peshenko, I.V.; Savchenko, A.B.; Makino, C.L.; Ho, Y.-S.; Fain, G.L.; Dizhoor, A.M. The Y99C Mutation in Guanylyl Cyclase-Activating Protein 1 Increases Intracellular Ca2+ and Causes Photoreceptor Degeneration in Transgenic Mice. J. Neurosci. 2004, 24, 6078–6085. [Google Scholar] [CrossRef] [Green Version]
  79. Peshenko, I.V.; Cideciyan, A.V.; Sumaroka, A.; Olshevskaya, E.V.; Scholten, A.; Abbas, S.; Koch, K.-W.; Jacobson, S.G.; Dizhoor, A.M. A G86R mutation in the calcium-sensor protein GCAP1 alters regulation of retinal guanylyl cyclase and causes dominant cone-rod degeneration. J. Biol. Chem. 2019, 294, 3476–3488. [Google Scholar] [CrossRef] [Green Version]
  80. Huang, S.H.; Pittler, S.J.; Huang, X.; Oliveira, L.; Berson, E.L.; Dryja, T.P. Autosomal recessive retinitis pigmentosa caused by mutations in the α subunit of rod cGMP phosphodiesterase. Nat. Genet. 1995, 11, 468–471. [Google Scholar] [CrossRef]
  81. McLaughlin, M.E.; Ehrhart, T.L.; Berson, E.L.; Dryja, T.P. Mutation spectrum of the gene encoding the beta subunit of rod phosphodiesterase among patients with autosomal recessive retinitis pigmentosa. Proc. Natl. Acad. Sci. USA 1995, 92, 3249–3253. [Google Scholar] [CrossRef]
  82. Danciger, M.; Blaney, J.; Gao, Y.; Zhao, D.; Heckenlively, J.R.; Jacobson, S.G.; Farber, D.B. Mutations in the PDE6B Gene in Autosomal Recessive Retinitis Pigmentosa. Genomics 1995, 30, 1–7. [Google Scholar] [CrossRef] [PubMed]
  83. Dryja, T.P.; Rucinski, D.E.; Chen, S.H.; Berson, E.L. Frequency of mutations in the gene encoding the alpha subunit of rod cGMP-phosphodiesterase in autosomal recessive retinitis pigmentosa. Investig. Ophthalmol. Vis. Sci. 1999, 40, 1859–1865. [Google Scholar]
  84. Gopalakrishna, K.N.; Boyd, K.; Artemyev, N.O. Mechanisms of mutant PDE6 proteins underlying retinal diseases. Cell Signal. 2017, 37, 74–80. [Google Scholar] [CrossRef]
  85. Kohl, S.; Coppieters, F.; Meire, F.; Schaich, S.; Roosing, S.; Brennenstuhl, C.; Bolz, S.; van Genderen, M.M.; Riemslag, F.C.; Lukowski, R.; et al. A Nonsense Mutation in PDE6H Causes Autosomal-Recessive Incomplete Achromatopsia. Am. J. Hum. Genet. 2012, 91, 527–532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Grau, T.; Artemyev, N.O.; Rosenberg, T.; Dollfus, H.; Haugen, O.H.; Sener, E.C.; Jurklies, B.; Andreasson, S.; Kernstock, C.; Larsen, M.; et al. Decreased catalytic activity and altered activation properties of PDE6C mutants associated with autosomal recessive achromatopsia. Hum. Mol. Genet. 2011, 20, 719–730. [Google Scholar] [CrossRef] [Green Version]
  87. Thiadens, A.A.; Hollander, A.I.D.; Roosing, S.; Nabuurs, S.B.; Zekveld-Vroon, R.C.; Collin, R.W.; De Baere, E.; Koenekoop, R.K.; van Schooneveld, M.J.; Strom, T.M.; et al. Homozygosity Mapping Reveals PDE6C Mutations in Patients with Early-Onset Cone Photoreceptor Disorders. Am. J. Hum. Genet. 2009, 85, 240–247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Chang, B.; Grau, T.; Dangel, S.; Hurd, R.; Jurklies, B.; Sener, E.C.; Andreasson, S.; Dollfus, H.; Baumann, B.; Bolz, S.; et al. A homologous genetic basis of the murine cpfl1 mutant and human achromatopsia linked to mutations in the PDE6C gene. Proc. Natl. Acad. Sci. USA 2009, 106, 19581–19586. [Google Scholar] [CrossRef]
  89. Farber, D.; Lolley, R.N. Enzymic basis for cyclic GMP accumulation in degenerative photoreceptor cells of mouse retina. J. Cycl. Nucleotide Res. 1976, 2, 139–148. [Google Scholar]
  90. Farber, D.B. From mice to men: The cyclic GMP phosphodiesterase gene in vision and disease. The Proctor Lecture. Investig. Ophthalmol. Vis. Sci. 1995, 36, 263–275. [Google Scholar]
  91. Farber, D.B.; Lolley, R.N. Cyclic Guanosine Monophosphate: Elevation in Degenerating Photoreceptor Cells of the C3H Mouse Retina. Science 1974, 186, 449–451. [Google Scholar] [CrossRef]
  92. Portera-Cailliau, C.; Sung, C.H.; Nathans, J.; Adler, R. Apoptotic photoreceptor cell death in mouse models of retinitis pigmentosa. Proc. Natl. Acad. Sci. USA 1994, 91, 974–978. [Google Scholar] [CrossRef]
  93. Tsang, S.H.; Gouras, P.; Yamashita, C.K.; Kjeldbye, H.; Fisher, J.; Farber, D.B.; Goff, S.P. Retinal degeneration in mice lacking the gamma subunit of the rod cGMP phosphodiesterase. Science 1996, 272, 1026–1029. [Google Scholar] [CrossRef]
  94. Chang, B.; Hawes, N.; Hurd, R.; Davisson, M.; Nusinowitz, S.; Heckenlively, J. Retinal degeneration mutants in the mouse. Vis. Res. 2002, 42, 517–525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Hasegawa, T.; Ikeda, H.O.; Nakano, N.; Muraoka, Y.; Tsuruyama, T.; Okamoto-Furuta, K.; Kohda, H.; Yoshimura, N. Changes in morphology and visual function over time in mouse models of retinal degeneration: An SD-OCT, histology, and electroretinography study. Jpn. J. Ophthalmol. 2016, 60, 111–125. [Google Scholar] [CrossRef] [PubMed]
  96. Barhoum, R.; Martínez-Navarrete, G.; Corrochano, S.; Germain, F.; Fernandez-Sanchez, L.; de la Rosa, E.; de la Villa, P.; Cuenca, N. Functional and structural modifications during retinal degeneration in the rd10 mouse. Neuroscience 2008, 155, 698–713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Wang, T.; Reingruber, J.; Woodruff, M.L.; Majumder, A.; Camarena, A.; Artemyev, N.; Fain, G.; Chen, J. The PDE6 mutation in the rd10 retinal degeneration mouse model causes protein mislocalization and instability and promotes cell death through increased ion influx. J. Biol. Chem. 2018, 293, 15332–15346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Trifunović, D.; Dengler, K.; Michalakis, S.; Zrenner, E.; Wissinger, B.; Paquet-Durand, F. cGMP-dependent cone photoreceptor degeneration in the cpfl1 mouse retina. J. Comp. Neurol. 2010, 518, 3604–3617. [Google Scholar] [CrossRef]
  99. Schaeferhoff, K.; Michalakis, S.; Tanimoto, N.; Fischer, M.D.; Becirovic, E.; Beck, S.C.; Huber, G.; Rieger, N.; Riess, O.; Wis-singer, B.; et al. Induction of STAT3-related genes in fast degenerating cone photo-receptors of cpfl1 mice. Cell Mol. Life Sci. 2010, 67, 3173–3186. [Google Scholar] [CrossRef]
  100. Fischer, M.D.; Tanimoto, N.; Beck, S.C.; Huber, G.; Schaeferhoff, K.; Michalakis, S.; Riess, O.; Wissinger, B.; Biel, M.; Bonin, M.; et al. Structural and Functional Phenotyping in the Cone-Specific Photoreceptor Function Loss 1 (cpfl1) Mouse Mutant—A Model of Cone Dystrophies. Adv. Exp. Med. Biol. 2010, 664, 593–599. [Google Scholar] [CrossRef] [Green Version]
  101. Kandaswamy, S.; Zobel, L.; John, B.; Santhiya, S.T.; Bogedein, J.; Przemeck, G.K.H.; Gailus-Durner, V.; Fuchs, H.; Biel, M.; de Angelis, M.H.; et al. Mutations within the cGMP-binding domain of CNGA1 causing autosomal recessive retinitis pigmentosa in human and animal model. Cell Death Discov. 2022, 8, 387. [Google Scholar] [CrossRef]
  102. Jin, X.; Qu, L.-H.; Hou, B.-K.; Xu, H.-W.; Meng, X.-H.; Pang, C.-P.; Yin, Z.-Q. Novel compound heterozygous mutation in the CNGA1 gene underlie autosomal recessive retinitis pigmentosa in a Chinese family. Biosci. Rep. 2016, 36, e00289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Radojevic, B.; Jones, K.; Klein, M.; Mauro-Herrera, M.; Kingsley, R.; Birch, D.G.; Bennett, L.D. Variable expressivity in patients with autosomal recessive retinitis pigmentosa associated with the gene CNGB1. Ophthalmic Genet. 2021, 42, 15–22. [Google Scholar] [CrossRef] [PubMed]
  104. Katagiri, S.; Akahori, M.; Sergeev, Y.; Yoshitake, K.; Ikeo, K.; Furuno, M.; Hayashi, T.; Kondo, M.; Ueno, S.; Tsunoda, K.; et al. Whole Exome Analysis Identifies Frequent CNGA1 Mutations in Japanese Population with Autosomal Recessive Retinitis Pigmentosa. PLoS ONE 2014, 9, e108721. [Google Scholar] [CrossRef] [PubMed]
  105. Saito, K.; Gotoh, N.; Kang, I.; Shimada, T.; Usui, T.; Terao, C. A case of retinitis pigmentosa homozygous for a rare CNGA1 causal variant. Sci. Rep. 2021, 11, 4681. [Google Scholar] [CrossRef]
  106. Nassisi, M.; Smirnov, V.; Hernandez, C.S.; Mohand-Saïd, S.; Condroyer, C.; Antonio, A.; Kühlewein, L.; Kempf, M.; Kohl, S.; Wissinger, B.; et al. CNGB1-related rod-cone dystrophy: A mutation review and update. Hum. Mutat. 2021, 42, 641–666. [Google Scholar] [CrossRef]
  107. Kohl, S.; Baumann, B.; Broghammer, M.; Jägle, H.; Sieving, P.; Kellner, U.; Spegal, R.; Anastasi, M.; Zrenner, E.; Sharpe, L.T.; et al. Mutations in the CNGB3 gene encoding the beta-subunit of the cone photoreceptor cGMP-gated channel are responsible for achromatopsia (ACHM3) linked to chromosome 8q21. Hum. Mol. Genet. 2000, 9, 2107–2116. [Google Scholar] [CrossRef]
  108. Wissinger, B.; Gamer, D.; Jägle, H.; Giorda, R.; Marx, T.; Mayer, S.; Tippmann, S.; Broghammer, M.; Jurklies, B.; Rosenberg, T.; et al. CNGA3 Mutations in Hereditary Cone Photoreceptor Disorders. Am. J. Hum. Genet. 2001, 69, 722–737. [Google Scholar] [CrossRef] [Green Version]
  109. Sidjanin, D.J.; Lowe, J.K.; McElwee, J.; Milne, B.S.; Phippen, T.M.; Sargan, D.R.; Aguirre, G.D.; Acland, G.M.; Ostrander, E. Canine CNGB3 mutations establish cone degeneration as orthologous to the human achromatopsia locus ACHM3. Hum. Mol. Genet. 2002, 11, 1823–1833. [Google Scholar] [CrossRef] [Green Version]
  110. Kohl, S.; Varsanyi, B.; Antunes, G.A.; Baumann, B.; Hoyng, C.B.; Jägle, H.; Rosenberg, T.; Kellner, U.; Lorenz, B.; Salati, R.; et al. CNGB3 mutations account for 50% of all cases with autosomal recessive achromatopsia. Eur. J. Hum. Genet. 2005, 13, 302–308. [Google Scholar] [CrossRef] [Green Version]
  111. Nishiguchi, K.M.; Sandberg, M.A.; Gorji, N.; Berson, E.L.; Dryja, T.P. Cone cGMP-gated channel mutations and clinical findings in patients with achromatopsia, macular degeneration, and other hereditary cone diseases. Hum. Mutat. 2005, 25, 248–258. [Google Scholar] [CrossRef]
  112. Liu, Y.; Wang, Y.; Xiao, Y.; Li, X.; Ruan, S.; Luo, X.; Wan, X.; Wang, F.; Sun, X. Retinal degeneration in mice lacking the cyclic nucleotide-gated channel subunit CNGA1. FASEB J. 2021, 35, e21859. [Google Scholar] [CrossRef] [PubMed]
  113. Xu, J.; Morris, L.; Thapa, A.; Ma, H.; Michalakis, S.; Biel, M.; Baehr, W.; Peshenko, I.V.; Dizhoor, A.M.; Ding, X.Q. cGMP accumulation causes photoreceptor degeneration in CNG channel deficiency: Evidence of cGMP cytotoxicity independently of enhanced CNG channel function. J. Neurosci. 2013, 33, 14939–14948. [Google Scholar] [CrossRef] [Green Version]
  114. Thapa, A.; Morris, L.; Xu, J.; Ma, H.; Michalakis, S.; Biel, M.; Ding, X.-Q. Endoplasmic Reticulum Stress-associated Cone Photoreceptor Degeneration in Cyclic Nucleotide-gated Channel Deficiency. J. Biol. Chem. 2012, 287, 18018–18029. [Google Scholar] [CrossRef] [Green Version]
  115. Michalakis, S.; Xu, J.; Biel, M.; Ding, X.-Q. Detection of cGMP in the Degenerating Retina. Methods Mol. Biol. 2013, 1020, 235–245. [Google Scholar] [CrossRef]
  116. Michalakis, S.; Muhlfriedel, R.; Tanimoto, N.; Krishnamoorthy, V.; Koch, S.; Fischer, M.D.; Becirovic, E.; Bai, L.; Huber, G.; Beck, S.C.; et al. Restoration of cone vision in the CNGA3-/- mouse model of congenital complete lack of cone photoreceptor function. Mol. Ther. 2010, 18, 2057–2063. [Google Scholar] [CrossRef]
  117. Wang, T.; Tsang, S.H.; Chen, J. Two pathways of rod photoreceptor cell death induced by elevated cGMP. Hum. Mol. Genet. 2017, 26, 2299–2306. [Google Scholar] [CrossRef] [Green Version]
  118. Paquet-Durand, F.; Bernhard-Kurz, S.; Arango-Gonzalez, B.; Zrenner, E.; Ueffing, M. Cell Death in rd2/rds Retina: An Apoptotic Process? Investig. Ophthalmol. Vis. Sci. 2012, 53, 6891. [Google Scholar]
  119. Ding, X.Q.; Nour, M.; Ritter, L.M.; Goldberg, A.F.; Fliesler, S.J.; Naash, M.I. The R172W mutation in peripherin/rds causes a cone-rod dystrophy in transgenic mice. Hum. Mol. Genet. 2004, 13, 2075–2087. [Google Scholar] [CrossRef]
  120. Stricker, H.M.; Ding, X.Q.; Quiambao, A.; Fliesler, S.J.; Naash, M.I. The Cys214-->Ser mutation in peripherin/rds causes a loss-of-function phenotype in transgenic mice. Biochem. J. 2005, 388, 605–613. [Google Scholar] [CrossRef] [Green Version]
  121. Tosi, J.; Davis, R.J.; Wang, N.-K.; Naumann, M.; Lin, C.-S.; Tsang, S.H. shRNA knockdown of guanylate cyclase 2e or cyclic nucleotide gated channel alpha 1 increases photoreceptor survival in a cGMP phosphodiesterase mouse model of retinitis pigmentosa. J. Cell Mol. Med. 2011, 15, 1778–1787. [Google Scholar] [CrossRef]
  122. Ma, H.; Butler, M.R.; Thapa, A.; Belcher, J.; Yang, F.; Baehr, W.; Biel, M.; Michalakis, S.; Ding, X.-Q. cGMP/Protein Kinase G Signaling Suppresses Inositol 1,4,5-Trisphosphate Receptor Phosphorylation and Promotes Endoplasmic Reticulum Stress in Photoreceptors of Cyclic Nucleotide-gated Channel-deficient Mice. J. Biol. Chem. 2015, 290, 20880–20892. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Butler, M.R.; Ma, H.; Yang, F.; Belcher, J.; Le, Y.-Z.; Mikoshiba, K.; Biel, M.; Michalakis, S.; Iuso, A.; Križaj, D.; et al. Endoplasmic reticulum (ER) Ca2+-channel activity contributes to ER stress and cone death in cyclic nucleotide-gated channel deficiency. J. Biol. Chem. 2017, 292, 11189–11205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Yang, P.; Lockard, R.; Titus, H.; Hiblar, J.; Weller, K.; Wafai, D.; Weleber, R.G.; Duvoisin, R.M.; Morgans, C.W.; Pennesi, M.E. Suppression of cGMP-Dependent Photoreceptor Cytotoxicity with Mycophenolate Is Neuroprotective in Murine Models of Retinitis Pigmentosa. Investig. Ophthalmol. Vis. Sci. 2020, 61, 25. [Google Scholar] [CrossRef]
  125. Peshenko, I.V.; Olshevskaya, E.V.; Dizhoor, A.M. Retinal degeneration-3 protein attenuates photoreceptor de-generation in transgenic mice expressing dominant mutation of human retinal guanylyl cyclase. J. Biol. Chem. 2021, 297, 101201. [Google Scholar] [CrossRef] [PubMed]
  126. Morris, L.; Ma, Z.; Thapa, A.; Ma, H.; Michalakis, S.; Biel, M.; Baehr, W.; Ding, X.Q. Exploration of the Mechanisms of Cone Photoreceptor Death in the Deficiency of Phosphodiesterase. Investig. Ophthalmol. Vis. Sci. 2013, 54, 5953. [Google Scholar]
  127. Paquet-Durand, F.; Beck, S.; Michalakis, S.; Goldmann, T.; Huber, G.; Mühlfriedel, R.; Trifunović, D.; Fischer, M.D.; Fahl, E.; Duetsch, G.; et al. A key role for cyclic nucleotide gated (CNG) channels in cGMP-related retinitis pigmentosa. Hum. Mol. Genet. 2011, 20, 941–947. [Google Scholar] [CrossRef] [Green Version]
  128. Hofmann, F.; Bernhard, D.; Lukowski, R.; Weinmeister, P. cGMP Regulated Protein Kinases (cGK). In Handbook of Experimental Pharmacology; Springer: Berlin/Heidelberg, Germany, 2009; pp. 137–162. [Google Scholar] [CrossRef]
  129. Hofmann, F. The Biology of Cyclic GMP-dependent Protein Kinases. J. Biol. Chem. 2005, 280, 1–4. [Google Scholar] [CrossRef] [Green Version]
  130. Vighi, E.; Trifunovic, D.; Veiga-Crespo, P.; Rentsch, A.; Hoffmann, D.; Sahaboglu, A.; Strasser, T.; Kulkarni, M.; Bertolotti, E.; van den Heuvel, A.; et al. Combination of cGMP analogue and drug delivery system provides functional protection in hereditary retinal degeneration. Proc. Natl. Acad. Sci. USA 2018, 115, E2997–E3006. [Google Scholar] [CrossRef] [Green Version]
  131. Koch, M.; Scheel, C.; Ma, H.; Yang, F.; Stadlmeier, M.; Glück, A.F.; Murenu, E.; Traube, F.R.; Carell, T.; Biel, M.; et al. The cGMP-Dependent Protein Kinase 2 Contributes to Cone Photoreceptor Degeneration in the Cnga3-Deficient Mouse Model of Achromatopsia. Int. J. Mol. Sci. 2020, 22, 52. [Google Scholar] [CrossRef]
  132. Szegezdi, E.; Logue, S.E.; Gorman, A.M.; Samali, A. Mediators of endoplasmic reticulum stress-induced apoptosis. EMBO Rep. 2006, 7, 880–885. [Google Scholar] [CrossRef] [Green Version]
  133. Kim, I.; Xu, W.; Reed, J.C. Cell death and endoplasmic reticulum stress: Disease relevance and therapeutic opportunities. Nat. Rev. Drug Discov. 2008, 7, 1013–1030. [Google Scholar] [CrossRef] [PubMed]
  134. Chan, P.; Stolz, J.; Kohl, S.; Chiang, W.-C.; Lin, J.H. Endoplasmic reticulum stress in human photoreceptor diseases. Brain Res. 2016, 1648, 538–541. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Ma, H.; Thapa, A.; Morris, L.M.; Michalakis, S.; Biel, M.; Frank, M.B.; Bebak, M.; Ding, X.-Q. Loss of cone cyclic nucleotide-gated channel leads to alterations in light response modulating system and cellular stress response pathways: A gene expression profiling study. Hum. Mol. Genet. 2013, 22, 3906–3919. [Google Scholar] [CrossRef] [Green Version]
  136. Nakagawa, T.; Yuan, J. Cross-Talk between Two Cysteine Protease Families: Activation of caspase-12 by calpain in apoptosis. J. Cell Biol. 2000, 150, 887–894. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Nakagawa, T.; Zhu, H.; Morishima, N.; Li, E.; Xu, J.; Yankner, B.A.; Yuan, J. Caspase-12 mediates endoplasmic-reticulum-specific apoptosis and cytotoxicity by amyloid-beta. Nature 2000, 403, 98–103. [Google Scholar] [CrossRef]
  138. Haidara, K.; Marion, M.; Gascon-Barré, M.; Denizeau, F.; Averill-Bates, D.A. Implication of caspases and subcellular compartments in tert-butylhydroperoxide induced apoptosis. Toxicol. Appl. Pharmacol. 2008, 229, 65–76. [Google Scholar] [CrossRef]
  139. Sanges, D.; Comitato, A.; Tammaro, R.; Marigo, V. Apoptosis in retinal degeneration involves cross-talk between apoptosis-inducing factor (AIF) and caspase-12 and is blocked by calpain inhibitors. Proc. Natl. Acad. Sci. USA 2006, 103, 17366–17371. [Google Scholar] [CrossRef]
  140. Shoshan-Barmatz, V.; Zakar, M.; Shmuelivich, F.; Nahon, E.; Vardi, N. Retina expresses a novel variant of the ryanodine receptor. Eur. J. Neurosci. 2007, 26, 3113–3125. [Google Scholar] [CrossRef]
  141. Yang, F.; Ma, H.; Butler, M.R.; Ding, X. Potential contribution of ryanodine receptor 2 upregulation to cGMP/PKG signaling-induced cone degeneration in cyclic nucleotide-gated channel deficiency. FASEB J. 2020, 34, 6335–6350. [Google Scholar] [CrossRef]
  142. Ma, H.; Yang, F.; Butler, M.R.; Rapp, J.; Le, Y.-Z.; Ding, X.-Q. Ryanodine Receptor 2 Contributes to Impaired Protein Localization in Cyclic Nucleotide-Gated Channel Deficiency. Eneuro 2019, 6, 0119–19. [Google Scholar] [CrossRef] [Green Version]
  143. Yang, F.; Ma, H.; Butler, M.R.; Ding, X.Q. Preservation of endoplasmic reticulum (ER) Ca2+ stores by deletion of inositol-1,4,5-trisphosphate receptor type 1 promotes ER retrotranslocation, proteostasis, and protein outer segment localization in cyclic nucleotide-gated channel-deficient cone photoreceptors. FASEB J. 2021, 35, e21579. [Google Scholar]
  144. Roy, A.; Tolone, A.; Hilhorst, R.; Groten, J.; Tomar, T.; Paquet-Durand, F. Kinase activity profiling identifies putative downstream targets of cGMP/PKG signaling in inherited retinal neurodegeneration. Cell Death Discov. 2022, 8, 93. [Google Scholar] [CrossRef] [PubMed]
  145. Zhou, J.; Rasmussen, M.; Ekstrom, P. cGMP-PKG dependent transcriptome in normal and degenerating retinas: Novel insights into the retinitis pigmentosa pathology. Exp. Eye Res. 2021, 212, 108752. [Google Scholar] [CrossRef] [PubMed]
  146. Vosler, P.S.; Sun, D.; Wang, S.; Gao, Y.; Kintner, D.B.; Signore, A.P.; Cao, G.; Chen, J. Calcium dysregulation induces apoptosis-inducing factor release: Cross-talk between PARP-1- and calpain- signaling pathways. Exp. Neurol. 2009, 218, 213–220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. McGinnis, K.M.; Gnegy, M.E.; Park, Y.H.; Mukerjee, N.; Wang, K.K. Procaspase-3 and poly(ADP)ribose polymerase (PARP) are calpain substrates. Biochem. Biophys. Res. Commun. 1999, 263, 94–99. [Google Scholar] [CrossRef] [Green Version]
  148. Paquet-Durand, F.; Azadi, S.; Hauck, S.M.; Ueffing, M.; van Veen, T.; Ekström, P. Calpain is activated in degenerating photoreceptors in the rd1 mouse. J. Neurochem. 2006, 96, 802–814. [Google Scholar] [CrossRef]
  149. Paquet-Durand, F.; Sanges, D.; McCall, J.; Silva, J.; Van Veen, T.; Marigo, V.; Ekström, P. Photoreceptor rescue and toxicity induced by different calpain inhibitors. J. Neurochem. 2010, 115, 930–940. [Google Scholar] [CrossRef]
  150. Power, M.J.; Rogerson, L.E.; Schubert, T.; Berens, P.; Euler, T.; Paquet-Durand, F. Systematic spatiotemporal mapping reveals divergent cell death pathways in three mouse models of hereditary retinal degeneration. J. Comp. Neurol. 2020, 528, 1113–1139. [Google Scholar] [CrossRef] [Green Version]
  151. Vu, J.T.; Wang, E.; Wu, J.; Sun, Y.J.; Velez, G.; Bassuk, A.G.; Lee, S.H.; Mahajan, V.B. Calpains as mechanistic drivers and therapeutic targets for ocular disease. Trends Mol. Med. 2022, 28, 644–661. [Google Scholar] [CrossRef]
  152. Comitato, A.; Sanges, D.; Rossi, A.; Humphries, M.M.; Marigo, V. Activation of Bax in Three Models of Retinitis Pigmentosa. Investig. Ophthalmol. Vis. Sci. 2014, 55, 3555–3561. [Google Scholar] [CrossRef] [Green Version]
  153. Oka, T.; Nakajima, T.; Tamada, Y.; Shearer, T.R.; Azuma, M. Contribution of calpains to photoreceptor cell death in N-methyl-N-nitrosourea-treated rats. Exp. Neurol. 2007, 204, 39–48. [Google Scholar] [CrossRef]
  154. Baehr, W.; Palczewski, K. Guanylate Cyclase-Activating Proteins and Retina Disease. Subcell. Biochem. 2007, 45, 71–91. [Google Scholar] [CrossRef]
  155. Plana-Bonamaisó, A.; López-Begines, S.; Andilla, J.; Fidalgo, M.J.; Loza-Alvarez, P.; Estanyol, J.M.; de la Villa, P.; Méndez, A. GCAP neuronal calcium sensor proteins mediate photoreceptor cell death in the rd3 mouse model of LCA12 congenital blindness by involving endoplasmic reticulum stress. Cell Death Dis. 2020, 11, 62. [Google Scholar] [CrossRef] [Green Version]
  156. Sancho-Pelluz, J.; Alavi, M.V.; Sahaboglu, A.; Kustermann, S.; Farinelli, P.; Azadi, S.; van Veen, T.; Romero, F.J.; Paquet-Durand, F.; Ekström, P. Excessive HDAC activation is critical for neurodegeneration in the rd1 mouse. Cell Death Dis. 2010, 1, e24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Trifunović, D.; Arango-Gonzalez, B.; Comitato, A.; Barth, M.; Del Amo, E.M.; Kulkarni, M.; Sahaboglu, A.; Hauck, S.M.; Urtti, A.; Arsenijevic, Y.; et al. HDAC inhibition in the cpfl1 mouse protects de-generating cone photoreceptors in vivo. Hum. Mol. Genet. 2016, 25, 4462–4472. [Google Scholar] [PubMed] [Green Version]
  158. Ekstrom, P.A.; Ueffing, M.; Zrenner, E.; Paquet-Durand, F. Novel in situ activity assays for the quantitative molecular analysis of neurodegenerative processes in the retina. Curr. Med. Chem. 2014, 21, 3478–3493. [Google Scholar] [CrossRef] [PubMed]
  159. Wang, X.; Ge, P. Parthanatos in the pathogenesis of nervous system diseases. Neuroscience 2020, 449, 241–250. [Google Scholar] [CrossRef] [PubMed]
  160. Paquet-Durand, F.; Silva, J.; Talukdar, T.; Johnson, L.E.; Azadi, S.; van Veen, T.; Ueffing, M.; Hauck, S.M.; Ekström, P.A.R. Excessive Activation of Poly(ADP-Ribose) Polymerase Contributes to Inherited Photoreceptor Degeneration in the Retinal Degeneration 1 Mouse. J. Neurosci. 2007, 27, 10311–10319. [Google Scholar] [CrossRef]
  161. Zhang, F.; Xie, R.; Munoz, F.M.; Lau, S.S.; Monks, T.J. PARP-1 hyperactivation and reciprocal elevations in intracellular Ca2+ during ROS-induced nonapoptotic cell death. Toxicol. Sci. 2014, 140, 118–134. [Google Scholar] [CrossRef]
  162. Geistrikh, I.; Visochek, L.; Klein, R.; Miller, L.; Mittelman, L.; Shainberg, A.; Cohen-Armon, M. Ca2+-induced PARP-1 activation and ANF expression are coupled events in cardiomyocytes. Biochem. J. 2011, 438, 337–347. [Google Scholar] [CrossRef] [Green Version]
  163. Wenzel, A.; Grimm, C.; Samardzija, M.; Remé, C.E. Molecular mechanisms of light-induced photoreceptor apoptosis and neuroprotection for retinal degeneration. Prog. Retin. Eye Res. 2005, 24, 275–306. [Google Scholar] [CrossRef] [PubMed]
  164. Candé, C.; Vahsen, N.; Garrido, C.; Kroemer, G. Apoptosis-inducing factor (AIF): Caspase-independent after all. Cell Death Differ. 2004, 11, 591–595. [Google Scholar] [CrossRef] [PubMed]
  165. Wang, Y.; An, R.; Umanah, G.K.; Park, H.; Nambiar, K.; Eacker, S.M.; Kim, B.; Bao, L.; Harraz, M.M.; Chang, C.; et al. A nuclease that mediates cell death induced by DNA damage and poly(ADP-ribose) polymerase-1. Science 2016, 354, 6308. [Google Scholar] [CrossRef] [Green Version]
  166. Mizukoshi, S.; Nakazawa, M.; Sato, K.; Ozaki, T.; Metoki, T.; Ishiguro, S.-I. Activation of mitochondrial calpain and release of apoptosis-inducing factor from mitochondria in RCS rat retinal degeneration. Exp. Eye Res. 2010, 91, 353–361. [Google Scholar] [CrossRef]
  167. Comitato, A.; Schiroli, D.; Montanari, M.; Marigo, V. Calpain Activation Is the Major Cause of Cell Death in Photoreceptors Expressing a Rhodopsin Misfolding Mutation. Mol. Neurobiol. 2020, 57, 589–599. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The phototransduction and regulation of cGMP signaling in a representative rod photoreceptor. (A) In the dark, cGMP, at a high intracellular level, binds to the CNG channel to keep it open, which allows Ca2+ and Na+ influx and subsequently elevates [Ca2+]i. The Ca2+-bound GCAP binds and inactivates RetGC at high [Ca2+]i, exerting the negative feedback control to cGMP synthesis. In the meantime, Ca2+ ions are steadily extruded via the Na+/Ca2+-K+ exchanger (NCKX) in the outer segment and via Ca2+-ATPase from the inner segment (not shown), thus maintaining the constant free Ca2+ levels. (B) In light stimulation, photon (hv) absorption leads to a conformational rearrangement of the rhodopsin (R) protein, which activates multiple copies of the heterotrimeric G protein transducin (G), causing the exchange of GTP for GDP on its α subunit. The activated Gα subunit (GTP-bound) binds to the γ -subunit of PDE6, relieving the inhibitory constraint and leading to the catalytic acceleration of cGMP hydrolysis. Reduced free cGMP levels lead to the closure of the CNG channel, reduction in the Ca2+ influx, and the subsequent reduction in [Ca2+]i. Along with the continued activity of NCKX, the membrane is hyperpolarized, the electro-chemical signal is transmitted, and the synaptic glutamate release is altered. As a result of reduced [Ca2+]i, the Mg2+-bound/Ca2+-free GCAP binds to RetGC and activates cGMP synthesis, forming a feedback loop to open the CNG channel over again (R, rhodopsin; G, transducin; CNGC, CNG channel; NCKX, Na+/Ca2+-K+ exchanger; modified from Leskov et al. [7] and Tolone et al. [13].
Figure 1. The phototransduction and regulation of cGMP signaling in a representative rod photoreceptor. (A) In the dark, cGMP, at a high intracellular level, binds to the CNG channel to keep it open, which allows Ca2+ and Na+ influx and subsequently elevates [Ca2+]i. The Ca2+-bound GCAP binds and inactivates RetGC at high [Ca2+]i, exerting the negative feedback control to cGMP synthesis. In the meantime, Ca2+ ions are steadily extruded via the Na+/Ca2+-K+ exchanger (NCKX) in the outer segment and via Ca2+-ATPase from the inner segment (not shown), thus maintaining the constant free Ca2+ levels. (B) In light stimulation, photon (hv) absorption leads to a conformational rearrangement of the rhodopsin (R) protein, which activates multiple copies of the heterotrimeric G protein transducin (G), causing the exchange of GTP for GDP on its α subunit. The activated Gα subunit (GTP-bound) binds to the γ -subunit of PDE6, relieving the inhibitory constraint and leading to the catalytic acceleration of cGMP hydrolysis. Reduced free cGMP levels lead to the closure of the CNG channel, reduction in the Ca2+ influx, and the subsequent reduction in [Ca2+]i. Along with the continued activity of NCKX, the membrane is hyperpolarized, the electro-chemical signal is transmitted, and the synaptic glutamate release is altered. As a result of reduced [Ca2+]i, the Mg2+-bound/Ca2+-free GCAP binds to RetGC and activates cGMP synthesis, forming a feedback loop to open the CNG channel over again (R, rhodopsin; G, transducin; CNGC, CNG channel; NCKX, Na+/Ca2+-K+ exchanger; modified from Leskov et al. [7] and Tolone et al. [13].
Ijms 24 11200 g001
Figure 2. The cellular and molecular mechanisms underlying cGMP signaling-induced photoreceptor degeneration. Overactivation of RetGC/GCAP or deficiency of PDE6 leads to the accumulation of cGMP and the subsequent intracellular Ca2+ overload. Deficiency of RD3 leads to the aberrant activation of the RetGC/GCAP complex in the inner segment, leading to excessive cGMP production. Deficiency of the CNG channel leads to a reduced intracellular Ca2+ level and the subsequent accumulation of cGMP via the Ca2+/Mg2+-GCAP/RetGC complex. The deficiency of RetGC/GCAP leads to reduced cellular cGMP and [Ca2+]i. The elevated cGMP/PKG signaling and impaired Ca2+ homeostasis lead to cellular stress/death. The elevated cGMP/PKG signaling induces ER stress, leading to cell death via the activation of CHOP and caspase-7/12. Along with a reduced cellular Ca2+ level, it causes ER Ca2+ dysregulation by activating the IP3R1 and RyR2 channels and promoting the release of Ca2+ from the ER. The activated cGMP/PKG signaling may also induce HDAC and PARP, leading to cellular stress/cell death, directly and via the release of AIF from the mitochondria. The elevated cGMP/Ca2+/calpain signaling induces cellular stress/death by activating caspases, releasing AIF from the mitochondria, and stimulating PARP.
Figure 2. The cellular and molecular mechanisms underlying cGMP signaling-induced photoreceptor degeneration. Overactivation of RetGC/GCAP or deficiency of PDE6 leads to the accumulation of cGMP and the subsequent intracellular Ca2+ overload. Deficiency of RD3 leads to the aberrant activation of the RetGC/GCAP complex in the inner segment, leading to excessive cGMP production. Deficiency of the CNG channel leads to a reduced intracellular Ca2+ level and the subsequent accumulation of cGMP via the Ca2+/Mg2+-GCAP/RetGC complex. The deficiency of RetGC/GCAP leads to reduced cellular cGMP and [Ca2+]i. The elevated cGMP/PKG signaling and impaired Ca2+ homeostasis lead to cellular stress/death. The elevated cGMP/PKG signaling induces ER stress, leading to cell death via the activation of CHOP and caspase-7/12. Along with a reduced cellular Ca2+ level, it causes ER Ca2+ dysregulation by activating the IP3R1 and RyR2 channels and promoting the release of Ca2+ from the ER. The activated cGMP/PKG signaling may also induce HDAC and PARP, leading to cellular stress/cell death, directly and via the release of AIF from the mitochondria. The elevated cGMP/Ca2+/calpain signaling induces cellular stress/death by activating caspases, releasing AIF from the mitochondria, and stimulating PARP.
Ijms 24 11200 g002
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, S.; Ma, H.; Yang, F.; Ding, X. cGMP Signaling in Photoreceptor Degeneration. Int. J. Mol. Sci. 2023, 24, 11200. https://doi.org/10.3390/ijms241311200

AMA Style

Li S, Ma H, Yang F, Ding X. cGMP Signaling in Photoreceptor Degeneration. International Journal of Molecular Sciences. 2023; 24(13):11200. https://doi.org/10.3390/ijms241311200

Chicago/Turabian Style

Li, Shujuan, Hongwei Ma, Fan Yang, and Xiqin Ding. 2023. "cGMP Signaling in Photoreceptor Degeneration" International Journal of Molecular Sciences 24, no. 13: 11200. https://doi.org/10.3390/ijms241311200

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

Article Metrics

Back to TopTop