IMR Press / FBL / Volume 28 / Issue 10 / DOI: 10.31083/j.fbl2810240
Open Access Review
The Physiopathologic Roles of Calcium Signaling in Podocytes
Show Less
1 Department of Nephrology, Wuhan Union Hospital, Tongji Medical College, Huazhong University of Science and Technology, 430000 Wuhan, Hubei, China
*Correspondence: drylj@hotmail.com (Li-Jun Yao)
Front. Biosci. (Landmark Ed) 2023, 28(10), 240; https://doi.org/10.31083/j.fbl2810240
Submitted: 10 April 2023 | Revised: 18 May 2023 | Accepted: 8 June 2023 | Published: 18 October 2023
Copyright: © 2023 The Author(s). Published by IMR Press.
This is an open access article under the CC BY 4.0 license.
Abstract

Calcium (Ca2+) plays a critical role in podocyte function. The Ca2+-sensitive receptors on the cell surface can sense changes in Ca2+ concentration, and Ca2+ flow into podocytes, after activation of Ca2+ channels (such as transient receptor potential canonical (TRPC) channels and N-type calcium channels) by different stimuli. In addition, the type 2 ryanodine receptor (RyR2) and the voltage-dependent anion channel 1 (VDAC1) on mitochondrial store-operated calcium channels (SOCs) on the endoplasmic reticulum maintain the Ca2+ homeostasis of the organelle. Ca2+ signaling is transmitted through multiple downstream signaling pathways and participates in the morphogenesis, structural maintenance, and survival of podocytes. When Ca2+ is dysregulated, it leads to the occurrence and progression of various diseases, such as focal segmental glomerulosclerosis, diabetic kidney disease, lupus nephritis, transplant glomerulopathy, and hypertensive renal injury. Ca2+ signaling is a promising therapeutic target for podocyte-related diseases. This review first summarizes the role of Ca2+ sensing, Ca2+ channels, and different Ca2+-signaling pathways in the biological functions of podocytes, then, explores the status of Ca2+ signaling in different podocyte-related diseases and its advances as a therapeutic target.

Keywords
calcium
calcium signaling
disease
podocyte
target
1. Introduction

Podocytes are terminally differentiated epithelial cells located on the medial side of the Bowman’s capsule [1]. The nucleus of a podocyte is located in the center of the cell, which has a large cell body. Foot processes (FPs) are irregular protrusions extending from the cell body [2]. The basal area of the FP adheres to the glomerular basement membrane (GBM). In addition, adhesion molecules and connexins exist between individual FPs, thereby forming a porous structure known as the slit diaphragm (SD) [3]. The unique morphological and structural characteristics of podocytes are essential for the maintenance of the glomerular filtration barrier function. In a variety of pathologic conditions, the structure and morphology of podocytes are affected, and the glomerular filtration barrier is impaired, leading to the occurrence of podocyte-related diseases [4].

Acting as an important second messenger for cells, calcium (Ca2+) participates in and regulates various cellular activities, including cell apoptosis and proliferation [5]. Cellular and extracellular Ca2+ are in dynamic balance, and calcium-sensing receptors (CaSRs) can sense and regulate changes in Ca2+ concentrations [6]. At the same time, the strict regulation of intracellular Ca2+ is inseparable from the expression and activity of various channel proteins. Ca2+ signals transmitted by channel proteins are essential for maintaining podocyte function [7]. Ca2+ signaling is required for podocyte morphogenesis and FP formation [8]. Normal Ca2+ signaling can maintain the structure of podocytes by stabilizing cytoskeletal proteins [9, 10]. In addition, downstream signaling mediated by Ca2+ can regulate the autophagy balance in podocytes and participate in cell survival [11]. The function of podocytes is critically dependent on the processing of intracellular Ca2+ signaling.

In the pathological states of the podocyte, enhanced Ca2+ signaling in podocytes is observed [12]. Changes in Ca2+ transport and signaling are related to the occurrence and development of various podocyte-related diseases [9, 13, 14]. Further understanding of the status of Ca2+ signaling in podocytes could provide promising insights for the treatment of podocyte-related diseases. This review summarizes the current studies on Ca2+ sensing, Ca2+ channels, and the Ca2+-signaling pathway in podocytes, as well as the status and targets of the Ca2+-signaling pathway in podocyte-related diseases.

2. Ca2+ Sensing and Transport in Podocytes
2.1 Ca2+ Sensing

In mammals, Ca2+ homeostasis is mainly mediated by CaSRs [6]. In 2008, Piecha et al. [15] first described the presence of CaSRs in podocytes. The podocyte-specific CaSR knockout model suggested that the CaSR represented an important receptor in the maintenance of glomerular filtration barrier function [16].

A CaSR is a G-protein-coupled receptor (GPCR), which encodes a gene that is located on chromosome 3 in humans. CaSRs can be activated by type I and type II agonists. Type I agonists include divalent ions, such as Ca2+ and magnesium, β-amyloid peptide, and polycation, which directly stimulate the CaSR. Type II agonists are a class of regulatory substances, including spermidine, spermine, aromatic amino acid residues, extracellular pH, and ionic strength. In the presence of extracellular Ca2+, type II agonists increase the Ca2+ affinity of the CaSR by allostery [6]. The CaSR activates the downstream signaling cascade through coupling with heterotrimeric guanine nucleotide-binding proteins (G proteins), such as the Gi/o, G12/13, and Gq/11 families [17]. An increase in Ca2+ stimulates the activity of Gq/11 and phospholipase C (PLC), inducing the accumulation of inositol 1,4,5 triphosphates (IP3). Subsequently, IP3 attaches to the endoplasmic reticulum (ER), via the IP3 receptor, and induces Ca2+ influx via the storage manipulation channel located on the plasma membrane, which results in Ca2+ being released from the intracellular stores [18]. In podocytes, the activated CaSR can promote TRPC6-dependent Ca2+ entry [19].

The CaSR-mediated Ca2+ balance of podocytes, which is key for podocyte normal function, has gradually become a research focus.

2.2 Ca2+ Channels

The transport of Ca2+ in podocytes is associated with multiple channel proteins (Table 1, Ref. [20, 21, 22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32]). TRPCs, which belong to the transient receptor potential (TRP) superfamily, are the most widely studied channel family in podocytes [5]. There are seven isoforms of the TRPC proteins, including TRPV1-7. TRPCs are “non-selective” cation channels that can promote Ca2+ influx into the cells [33]. Studies demonstrated that TRPC1, TRPC3, TRPC4, TRPC5, and TRPC6 are all expressed in podocytes [7, 20, 21, 22], although only TPRC 3, TRPC5, and TRPC6 were shown to contribute to Ca2+ entry into the podocyte [23]. In podocytes, TRPCs can be activated by a variety of upstream signals, whereas Ca2+ stimulates the downstream Ca2+-dependent signaling cascade [34, 35, 36]. Depending on their location in the cell, TRPCs are involved in different cellular physiological processes. In organelles such as the ER, Golgi, and lysosomes, their activation also promotes Ca2+ influx [24].

Table 1.Subcellular localization and function of Ca2+ channels in podocytes.
Calcium channels Subcellular localization Function Reference
TRPCs TRPC1 Unclear Promotes Ca2+ influx [20]
TRPC4
TRPC3 Plasma membrane and organelle membranes, such as ER, Golgi, lysosomes, etc. Promotes Ca2+ influx [21, 24, 31]
TRPC5 [22, 24, 32]
TRPC6 [22, 23, 24]
SOCs STIM1 ER Sensor for luminal Ca2+ concentration in ER [25, 26, 27]
ORAI1 Plasma membrane Promotes an influx of Ca2+ into the cytoplasm
RyR 2 ER Calcium release from ER [28]
VDAC 1 Mitochondria Transports Ca2+ into the mitochondria [29]
Cav2.2 Plasma membrane Promotes Ca2+ influx [30]

Abbreviations: TRPCs, transient receptor potential canonical channels; SOCs, store-operated calcium channels; STIM1, stromal-interacting molecule 1; ORAI1, Ca2+-release-activated Ca2+ channel 1; RyR 2, type 2 ryanodine receptor; VDAC 1, voltage-dependent anion channel 1; Cav2.2, N-type calcium channel; ER, endoplasmic reticulum.

SOCs in the plasma membrane are triggered and activated by the depletion of Ca2+ from the ER. SOCs consist of a single-channel transmembrane protein, STIM1, in the ER, and a small plasma membrane protein, ORAI1 [25, 26]. When Ca2+ is depleted in the ER, STIM1 is activated and aggregates in the vicinity of ER-plasma membrane junctions. STIM1 activates ORAI1, resulting in an influx of Ca2+ into the cytoplasm [27]. The expressions of STIM1 and ORAI1 were observed in mouse podocytes [37]. In addition, overexpression of STIM1 and ORAI1 increased the Ca2+ influx in podocytes [37].

Increased intracellular Ca2+ is a burden on organelles that may cause dysfunction [38]. RyR2 in the ER, as well as VDAC1 in the mitochondria, mediate Ca2+ transport in the organelles [28, 29]. Researchers have mentioned the role of the RyR2/Ca2+-release channel in ER–Ca2+ balance [28]. In podocytes, VDAC1 controls the Ca2+ transport into and out of the mitochondria, which is essential for mitochondrial function [29]. The N-type calcium channel, Cav2.2, in podocytes, as a Ca2+ transport channel, may be associated with the activation of the local renin-angiotensin system (RAS) [30].

These Ca2+ transport channels are important components of cells and organelles. Due to the different pathophysiological statuses, activity changes in these channels cause an imbalance in cellular Ca2+ homeostasis, which may relate to the pathogenesis of some podocyte-related diseases [39]. Thus, using calcium channels as therapeutic targets has the potential for podocyte-related diseases.

3. Role of Ca2+ Signaling in Podocyte Cell Biology and Physiology

Ca2+ signaling is necessary for multiple pathophysiological functions of podocytes. Recent studies have shown that Ca2+ signaling is involved in the morphology, architecture, and survival of podocytes. We now describe in detail the roles of Ca2+ signaling in different biological processes involving podocytes.

3.1 Podocyte Morphology

Podocytes are essential for the maintenance of the glomerular filtration barrier. Studies have shown that mutations in TRPC6 associated with nephrotic syndrome have led to the disruption of the filtration barrier through elevated cytoplasmic Ca2+ in podocytes [40, 41]. The distribution of TRPC6 channels changes as podocyte differentiation matures [42]. This suggests that Ca2+ signaling may play a critical role in filtration-barrier formation. A study used transgenic zebrafish larvae and human kidney organoids to explore the role of Ca2+ signaling during podocyte development [8]. They found that Ca2+ signaling was active during podocyte differentiation, resulting from the triggering and release of intracellular Ca2+. This Ca2+ influx may be mediated through PLCg/IP3, leading to podocyte-foot-process differentiation and glomerular-filtration-barrier formation. The role of Ca2+ signaling in podocyte morphogenesis cannot be ignored and deserves more attention.

3.2 Podocyte Architecture

The tight and complex interactions between the slit diaphragm, focal adhesion, and the actin cytoskeleton are essential for maintaining the normal structure and function of podocytes [43]. Slit diaphragm proteins, focal adhesion protein complexes, and actin cytoskeleton proteins play important roles in the precise organization and regulatory role of the actin cytoskeleton. Ca2+ signaling can have an impact on podocyte structure by affecting cytoskeletal proteins.

Podocin and nephrin are known to be critical SD proteins in maintaining podocyte structural stability [44]. Lu et al. [10] elucidated the role of TRPC6 in high glucose-induced podocin reduction. In their study, TRPC6 knockdown resulted in a diminution of the decreased podocin in podocytes, and a high level of glucose decreased podocin through activating TRPC6. This suggested that podocin can be mediated by TRPC6, the activation of which results in an increase in intracellular Ca2+. However, the exact mechanism of action between TRPC6 and podocin remains unclear. Ca2+ is known to regulate protein degradation through ubiquitination [45], which could be an interesting focus for future research.

Bhargava et al. [9] reported that IgG from kidney transplant recipients with transplant glomerulopathy activated the calmodulin kinase IV (CaMK4)/glycogen synthase kinase-3 β (GSK3β) pathway. Phosphorylated GSK 3β stabilized nephrin transcriptional repressor SNAIL, which caused a reduction in nephrin expression [9]. Their group also demonstrated that a reduction in nephrin expression, induced by IgG from lupus nephritis, was associated with the activation of CaMK4 [46]. Although Ca2+ channels were not mentioned in these studies, IgG, which has been produced in the pathological state, has been shown to trigger TRPCs in several studies [47, 48]. This may also be the mechanism through which IgG activates CaMK4 in kidney diseases.

Several signaling pathways, including CaMK4, play essential roles in actin remodeling after an increase in Ca2+. Synaptopodin, an actin-binding protein, regulates the integrity and motility of podocytes through the Rho family of small GTPases (including Rac1, Cdc42, and RhoA) [49]. The pCaMK4 phosphorylates scaffold protein 14-3-3, which leads to the release and degradation of synaptic foot proteins [50], by enhancing Rac1 expression and decreasing RhoA expression [50, 51]. Rac1 promotes cell motility by forming lamellipodia, and RhoA inhibits cell motility by promoting the formation of contractile actin [52]. Insulin triggers store-operated Ca2+ entry, which leads to a decrease in synaptopodin via the Ca2+/calcineurin pathway [9, 10, 53].

Talin-1 is a podocyte cytoskeleton-associated anchor protein and a specific target of calpain [54]. Studies have shown that aberrant Ca2+ signaling can be transmitted through the TRPC to activate downstream calpain and calcineurin, which then, exert effects on the actin cytoskeleton by reducing Talin-1 expression [43, 55]. Farmer et al. [55] reported that TRPC6 can interact with calpain 1 and calpain 2 and maintain calpain localization at the membrane. This physical interaction is significant for cytoskeletal stability in podocytes, where the actin cytoskeleton forms a highly dynamic contractile state, which is critical for the integrity of the foot processes in podocytes [56]. The CaSR is directly connected to the cytoskeleton by binding to the thin filament protein A. Oh et al. [57] showed that CaSR activation increased the actin density and content in podocytes, and reduced the degradation of the actin cytoskeleton caused by PAN.

The biological aspects of Ca2+ signaling in podocytes cannot be ignored. Multiple pathological states cause podocyte structural changes via Ca2+ imbalance; thus, by regulating structure-related proteins, Ca2+ signaling also influences the mobility and stability of podocytes.

3.3 Podocyte Survival

The survival of podocytes is a complex physiological process associated with multiple factors, such as the integrity of the cytoskeleton and expression of various pro-survival genes, and autophagy [58, 59, 60]. The stability of the cytoskeleton underlies cell survival. Ca2+ signaling can affect podocyte survival by mediating the impairment of the cytoskeleton in podocytes. Recent studies have shown that Ca2+ signaling also regulates the expression of multiple survival factors. Oh et al. [57] reported that the CaSR mediates the survival of podocytes through the MAPK pathway. The transcription factor, the cAMP-response element-binding protein (CREB) is known to be a pro-survival gene activator, while BAD and Bcl-xl are also pro-survival factors [57]. When the CaSR was activated, the expression of phosphorylated CREB, BAD, and Bcl-xl increased. Hence, an activated CaSR mediates the phosphorylation of the p90 RSK through ERK1/2, ultimately leading to the phosphorylation of CREB, the pro-survival gene activator [57].

In addition to the CaSR, the relationship between downstream signaling mediated by Ca2+ channels and cell survival should not be ignored. Autophagy is essential for the homeostasis and survival of podocytes. Calpain in podocytes can be activated by Ca2+/calcineurin [54]. Bensaada et al. [61] found that the inhibition of calpain can maintain autophagy in podocytes during hypertension. The TRPC6 pathway also mediates autophagy in podocytes [62]. Ma et al. [11] reported that AngII can activate TRPC 6 and cause an increase in Ca2+ influx, which results in the activation of calmodulin-dependent protein kinase-β (CaMKKβ), which, when activated, promotes the formation of activated protein kinase (AMPK) and autophagy in podocytes. AngII-induced abnormal autophagy and damage in the podocyte can be alleviated following the inhibition of TRPC6-mediated Ca2+ influx or downstream Ca2+-signaling pathways, such as calcineurin [11, 63].

Stabilization of Ca2+ channels in multiple organelles is also essential for maintaining podocyte homeostasis. The mitochondrial Ca2+ unidirectional transporter (MCU) maintains mitochondrial Ca2+ homeostasis to ensure its normal function [64]. RyR2 in the ER is involved in ER–Ca2+ leakage as well as ER stress [28]. The trigger of SOCE in the ER also has a major impact on podocyte homeostasis [65].

4. Ca2+ Signaling in Podocyte-Related Diseases

As critical components of the glomerular filtration barrier, podocytes can modulate glomerular filtration function. Podocyte dysfunction is central to the pathophysiology of many common glomerular diseases, including diabetic kidney disease, glomerulonephritis, and genetic forms of nephrotic syndrome (NS) [66]. By intravital imaging of podocyte Ca2+, one study suggested that the conduction of Ca2+ signaling was a key pathogenic mechanism in various podocyte-related diseases [12]. Ca2+ signaling is involved in various podocyte-related kidney diseases, including focal segmental glomerulosclerosis, diabetic kidney disease, lupus nephritis, transplant glomerulopathy, and hypertensive renal injury. Below we describe the mechanisms of Ca2+ signaling in various diseases.

4.1 Focal Segmental Glomerulosclerosis (FSGS)

FSGS is an important cause of end-stage renal failure (ESRD), characterized by podocyte loss and dysfunction. Most patients show steroid-resistant nephrotic syndrome (SRNS).

Several studies have shown that genetic FSGS is associated with mutations in the TRPC6 gene [39, 67, 68, 69] (Fig. 1). Among the mutated genotypes, TRPC6 always showed a gained functional phenotype involving Ca2+-mediated podocyte pathogenic mechanisms [13, 70]. The mutant form of TRPC6 that causes FSGS also increases intracellular Ca2+. Multiple studies have shown that downstream Ca2+ signaling participates in podocyte damage in FSGS after the delayed inactivation of TRPC6. Ca2+ activates calcineurin [71] and CaMK4 [50], resulting in a decrease in nephrin and synaptopodin expressions, which are critical for podocyte structure. The gain-of-function TRPC6 mutant led to the phosphorylation of ERK1/2 in podocytes, which is involved in the regulation of cell proliferation and survival [72]. However, the pathogenic role of TRPC6 activation may be through activating NFAT-dependent transcription [73]. As substrates of calcineurin, activation of NFATc transcription factors leads to podocyte dysfunction, and ultimately, to glomerulosclerosis [73].

Fig. 1.

Calcium signaling in focal segmental glomerulosclerosis.

The loss-of-function TRPC6 mutation has been shown to be associated with the pathogenesis of FSGS, although the mechanism is currently unclear [67]. Farmer et al. [55] reported that TRPC6-/- podocytes showed decreased motility, stronger adhesion, and an altered actin cytoskeleton. FAK cleavage is involved in cell death and motility, whereas talin-1 cleavage is associated with cell-adhesion breakdown [74]. Calpain activity was lost and appeared to be mislocalized after the knockdown of TRPC6. In addition, cleavage of the calpain targets, talin-1 and FAK, was reduced [55], which may account for the pathogenesis of the TRPC6 loss-of-function mutation.

In addition to the functional changes of TRPC6 and downstream pathways contributing to podocyte damage in FSGS, RyR2 may also play a role in the pathogenesis of FSGS. In the NS/FSGS mouse model, the RyR2 channels in the ER of podocytes are phosphorylated, which leads to Ca2+ depletion and stress in the ER, resulting in podocyte damage [28].

4.2 Diabetic Kidney Disease (DKD)

An increase in Ca2+-channel activity was observed in the diabetic kidney disease model [14]; the knockdown of Ca2+ channels in Akita mice promoted insulin resistance and aggravated glomerular injury [75]. These results highlight the importance of the roles Ca2+ channels perform in the pathophysiological state of DKD.

As a metabolic-related kidney disease, the underlying pathogenesis of DKD, especially in the dysfunction of mitochondria, is gradually gaining attention [76]. Several studies have shown that Ca2+ signaling participates in the development of podocyte injury in DKD by regulating mitochondrial function (Fig. 2). Yu et al. [77] reported that TRPC6 mediated high-glucose-induced mitochondrial fission in DKD podocyte injury. TRPC6 mediates mitochondrial fission by Ca2+-activated calpain1 [77]. Tao et al. [78] found that an increase in SOCE in podocytes was induced by high glucose and angiotensin II (AngII), thereby mediating impaired mitochondrial membrane potential (MMP), ATP and mitochondrial superoxide production, and mitochondrial respiratory damage in DKD. Disruption of Ca2+ homeostasis in mitochondria leads to increased oxidative stress and cell apoptosis [79]. The connection between the “mitochondria-associated membrane” (MAM) and the ER can maintain Ca2+ homeostasis in the mitochondria [80]. Wei et al. [81] reported that the activation of TRPV1 attenuated mitochondrial dysfunction in podocytes caused by hyperglycemia. Transient Ca2+ influx, mediated by TRPV1, reduced the transcription of Fundc1, a key molecule involved in MAM formation, by activating 5 AMP-activated protein kinase (AMPK). Therefore, the activation of TRPV1 reduced MAM formation and the transport of Ca2+ from the ER to the mitochondria [81]. When the ER releases Ca2+ through the inositol 1,4,5,-triphosphate receptor (IP3R), Ca2+ enters the mitochondrial membrane space via VDAC1 and finally enters the mitochondrial matrix through the mitochondrial Ca2+ unidirectional transport protein (MCU) [82]. High glucose and AngII induced enhanced effects of IP3R-Grp75-VDAC1-MCU, which increased Ca2+ in the mitochondria as well as increased active caspase-3 [64]. This process participates in podocyte apoptosis in DKD.

Fig. 2.

Calcium signaling in diabetic kidney disease.

High glucose-induced apoptosis is one of the key factors contributing to podocyte injury in DKD [83]. Upregulation of Ca2+/calcineurin (CaN) signaling promotes podocyte apoptosis, which is a key mechanism in podocyte injury in DKD [84]. Podocyte apoptosis may also relate to the increased expression of p38, caused by TRPC6 knockout in Akita mice [75]. Tao et al. [85] also showed that a high glucose-induced increase of SOCE led to decreased expression of the nephrin protein by promoting calpain activation. In addition to the above Ca2+ channels, the N-type calcium channel is also involved in proteinuria production due to podocyte damage in DKD [86].

4.3 Lupus Nephritis (LN)

LN, as a chronic immune-complex-mediated renal lesion, is mainly characterized by renal glomerular impairment [87]. Abnormal IgG can be transported and deposited in podocytes via the neonatal Fc receptor (FcRn), causing podocytopathies [88]. These recent studies have indicated that the Ca2+/CaMK4 pathway plays an essential role in IgG-mediated podocyte injury (Fig. 3).

Fig. 3.

Calcium signaling in lupus nephritis.

The expression of Ca2+/CaMK4 is significantly elevated in podocytes of LN patients and lupus-prone mice [46]. Ichinose et al. [88] showed that pathogenic IgG upregulated CaMK4 and altered the expression of some genes associated with podocyte injury, including CD86 costimulation and skeleton-related genes, such as nephrin and synaptopodin. Bhargava et al. [46] also reported that CaMK4 upregulation, mediated by aberrantly glycosylated IgG, can phosphorylate NF-κB, which decreases the expression of nephrin by upregulating SNAIL. In addition, CaMK4 can also modulate podocyte motility in LN through the regulation of Rac1 and RhoA, while can also regulate the degradation of synaptopodin in LN via the phosphorylated scaffold protein 14-3-3β [50]. These studies suggest that Ca2+/CaMK4 may be a valuable target for the treatment of podocyte injury in LN.

4.4 Transplant Glomerulopathy (TG)

Chronic antibody-mediated immune rejection, ultimately, leads to renal graft failure. TG is the main histological feature. Podocyte shedding was closely associated with the occurrence of proteinuria and eGFR decrease in TG [89]. Both TG and LN are thought to be the result of antibody-mediated damage. Bhargava et al. [9] reported that N-glycosylated IgG in TG patients can also enter podocytes via FcRn. N-glycosylated IgG activates CaMK4, which phosphorylates GSK3β. Phosphorylated GSK3β reduces nephrin expression, resulting in the regulation of podocyte migration and the rearrangement of the actin cytoskeleton [9] (Fig. 4).

Fig. 4.

Calcium signaling in transplant glomerulopathy.

4.5 Hypertensive Renal Injury

Hypertension damages podocytes due to increased pressure in the glomerular capillaries, leading to the loss of podocytes [90]. The activity of RAS is enhanced in hypertensive patients [91]. AngII-induced autophagy in the podocytes of spontaneously hypertensive rats was explored by Ma et al. [11]. They found that the activation of TRPC6 can be stimulated by AngII, thereby resulting in Ca2+ influx and CaMKKβ-AMPK pathway activation, which leads to podocyte autophagy. N-type calcium channel is also involved in AngII-induced podocyte injury [92], which may be related to the production of ROS caused by AngII [93]. Golosova et al. [94] showed that κ-opioid receptors (κORs)/TRPC6 signaling is activated in hypertensive rats and human podocytes, while pathological Ca2+ signaling leads to actin cytoskeleton rearrangement. These results suggest a critical role for Ca2+ signaling in hypertensive podocyte injury (Fig. 5).

Fig. 5.

Calcium signaling in hypertensive renal injury.

Ca2+ channels and Ca2+-signaling pathways are closely related to the pathogenesis and progression of podocyte-related diseases. Activation of Ca2+ channels, such as TRPC6, RyR2, SOCs, and VDAC1, has been reported to contribute to podocyte-related diseases. Ca2+/CaMK4 was found to be involved in podocyte injury in LN and TG, and the activation of TRPV1 attenuated mitochondrial dysfunction in podocytes in DKD. These results suggest that Ca2+ channels and Ca2+ signaling are promising targets for the treatment of podocyte-related diseases.

5. Targets of Ca2+-Signaling Pathways in Podocytes

The implementation of Ca2+-signaling pathways as promising therapeutic targets is the focus of researchers; thus, we summarized the current Ca2+-signaling targets associated with podocytes in Table 2 (Ref. [11, 16, 28, 57, 71, 86, 93, 95, 96, 97, 98, 99, 100, 101, 102, 103, 104, 105, 106, 107, 108, 109, 110].

Table 2.Drugs and inhibitors targeting Ca2+ signaling pathways in podocytes.
Drugs and inhibitors Targets Mechanism Disease Reference
R-568 CaSR Activates the sensitivity of CaSR to calcium ions Glomerulosclerosis [57]
Cinacalcet NS [16]
Atractylodis rhizoma water extract (ARE) TRPC 6 Inhibits the activation of TRPC 6/p-CaMK4 signaling Fructose-induced kidney disease [95, 96]
Astragaloside IV (AS-IV) Inhibits the activation of TRPC 6 DKD [97]
Tetrandrine Inhibit TRPC 6 expression and downstream RhoA/Rock and calpain 1 signaling pathway NS [98, 99, 100]
Taurine (2-aminoethenonic acid) Suppresses TRPC 6 expression by upregulating H2S synthesis DKD [101]
Semi-synthetic compound larixyl carbamate (LC) Inhibits TRPC6 activity [102]
Qian Yang Yu Yin (QYYY) granule Inhibits TRPC6/CaMKKβ pathway Hypertensive kidney injury [11]
Metformin Inhibits TRPC 6 expression through AMPKα1 activation DKD [103]
Inhibits the ZEB 2/TRPC 6 pathway CKD [104]
Liraglutide Reduces TRPC6 expression DKD [105]
Sildenafil (Viagra) Increases cGMP levels by inhibiting PDE 5 and cGMP, to reduce TRPC 6 expression DKD [106]
Riociguat Activates sGC, causing an increase in cGMP production and inhibiting the TRPC 6 pathway Adriamycin-induced nephropathy [107]
FK506 Reduces TRPC 6 expression DKD [110]
K201 RyR 2 Blocks RyR 2-Ser2808 phosphorylation NS [28]
MANF Restores defective RyR 2 function
Cilnidipine Cav 2.2 Inhibits N-calcium channels DKD [86, 93]
microRNA-30 family members TRPC, calcineurin, NFATC3 Inhibits TRPC 6, PPP3CA and PPP3CB, PPP3R1 and NFATC3 FSGS [71]
CNI calcineurin Inhibits the calcineurin/NFAT/Angptl4 pathway NS [108, 109]

Abbreviations: CaSR, calcium-sensing receptor; NS, nephrotic syndrome; DKD, Diabetic Kidney Disease; FK506, tacrolimus; MANF, mesencephalic astrocyte-derived neurotrophic factor; RyR 2, type 2 ryanodine receptor; Cav 2.2, N-type calcium channel; FSGS, Focal Segmental Glomerulosclerosis; TRPC, transient receptor potential canonical; NFATC3, Nuclear factor of activated T cell transcription factors; PPP3CA/B, protein phosphatase 3 catalytic subunit alpha/beta; CNI, calcineurin inhibitors; NFAT, Nuclear factor of activated T cell.

5.1 CaSR

The specific knockdown of CaSR in podocytes disrupted the actin cytoskeleton and reduced cell attachment and migration speed, resulting in proteinuria. Calcimimetics allosterically activate the sensitivity of the CaSRs to Ca2+.

Oh et al. [57] found that R-568, a calcimimetic that activates the CaSR signal, stabilizes the podocyte cytoskeleton and enhances cell survival, thereby reducing aminonucleoside (PAN)-induced glomerulosclerosis. In addition to R-568, the calcimimetic cinacalcet has also been used in children with nephrotic syndrome to relieve proteinuria [16]. Huang et al. [111] proposed that the CaSR has important molecular associations in the pathophysiology of primary membranous nephropathy. Activation of CaSRs may cause podocyte damage through Ca2+ imbalance. Yadav et al. [112] suggested that the Ca2+ imbalance caused by CaSR activation may also be one of the possible mechanisms of renal dysfunction in sepsis. These results suggest that CaSR is also a potential therapeutic target in primary membranous nephropathy and cardio–renal syndrome, which still needs further exploration.

5.2 Ca2+ Channels

Therapies targeting Ca2+ channels in podocytes are receiving increased attention.

The TRPC6 channel acts as the most important link in podocyte Ca2+ signaling. Multiple herbal extracts and synthetics can act to treat podocyte injury by inhibiting the activity of TRPC6 and its downstream pathways. Studies have shown that atractylodin and Atractylodis rhizoma water extract (ARE) components can prevent fructose-induced glomerular damage. Atractylodin inhibits the activation of TRPC6/p-CaMK4 signaling in podocyte hypermotility under fructose conditions [95]. ARE also has a therapeutic effect by downregulating the TRPC6/Ca2+/CaMK4 pathway resulting from reducing hydrogen peroxide (H2O2) and malondialdehyde (MDA) levels in the glomeruli [96]. As Astragaloside IV (AS-IV), an active ingredient isolated from Astragalus membranaceous (Fisch). Bge, inhibited Ca2+ channel activity by targeting TRPC6. AS-IV can reduce the Ca2+ influx caused by palmitic acid, thereby alleviating the apoptosis of podocytes [97]. Tetrandrine, the main active ingredient isolated from the Chinese herbal radix Stephania tetrandra, has been used in a Chinese traditional medicine preparation called Fangji Huangqi Tang for the treatment of NS [98]. Studies have indicated that tetrandine can protect podocytes from intracellular Ca2+ influx by inhibiting TRPC6 overexpression and downstream RhoA/Rock signaling [99], or the calpain 1 signaling pathway [100]. Taurine (2-aminoethenonic acid) is a semi-essential amino acid, synthesized from cysteine, which inhibits high glucose-induced podocyte damage [101]. Taurine increases cystathionine-γ-lyase (CSE) expression and suppresses TRPC6 expression by upregulating H2S synthesis, thereby alleviating the pathogenic effects of high glucose. Researchers composed a semi-synthetic compound larixyl carbamate (LC), extracted from larixol, as an inhibitor of TRPC 6 channel activity [102]. In addition to synthetic chemicals, Qian Yang Yu Yin (QYYY) granule, which is widely used in hypertensive kidney injury, has been shown to protect podocytes by inhibiting the TRPC6/CaMKKβ pathway [11]. The inhibitory effect of metformin on the TRPC6 pathway has also been demonstrated. Szrejder et al. [103] showed that metformin inhibited high-glucose-induced TRPC6 expression in podocytes by activating AMPKα1, thereby protecting the podocyte cytoskeleton. Metformin also abrogates hypoxia-induced podocyte injury by inhibiting the ZEB2/TRPC6 pathway [104]. Liraglutide, a GLP1 receptor agonist, reduced TRPC6 expression in the kidneys of Type 1 diabetic rats, thereby preventing the progression of diabetic kidney disease [105]. Previous drugs used in non-renal diseases were also revived by researchers. Sonneveld et al. [106] found that sildenafil (Viagra) increased cyclic guanosine phosphate (cGMP) levels by inhibiting phosphodiesterase type 5 (PDE5)- and cGMP-reduced TRPC6 expression and activity, to prevent the progression of glomerular damage. Hart et al. [107] also found that riociguat, through activating soluble guanylate cyclase (sGC), led to an increase in cGMP production, thereby inhibiting the TRPC6 pathway and podocyte damage caused by TRPC6-mediated Ca2+ influx. Wang et al. [113] reported that an α1-AR agonist (phenylephrine hydrochloride) could participate in the loss of the cytoskeletal structure by inducing a TRPC6-dependent increase in intracellular Ca2+ in human podocytes. This provides a possible rationale for the inhibition of TRPC6 by the α1-AR inhibitor (silodosin) [114], to improve podocyte injury.

The Ca2+-release channel RyR2 on the ER of the podocyte is phosphorylated in the NS/FSGS mouse model, causing an imbalance in intracellular Ca2+ homeostasis and ER stress. Park et al. [28] proposed that a novel compound, K201, blocked RyR2-Ser2808 phosphorylation and that a brain astrocyte-derived neurotrophic factor (MANF) restored the defective RyR2 function.

Albuminuria and ultrafiltration were significantly improved after specific blockade of the α-1 subunit of the N-type calcium channel Cav 2.2 in diabetic mice [86]. This suggests that N-type calcium channels are a future therapeutic target for podocyte injury. Cilnidipine is an L/N-type dihydropyridine calcium channel blocker (CCB), which can pharmacologically inhibit the N- type calcium channels [115]. Cilnidipine reduced the urinary protein/creatinine ratio in hypertensive patients with chronic kidney disease more effectively than amlodipine (L-type CCB) [116, 117]. A subsequent study suggested that it may reduce the AngII-induced damage through the inhibition of N-type calcium channels in podocytes [93].

5.3 Ca2+/Calcineurin Signaling

Presently, studies targeting Ca2+/calcineurin are of great interest to researchers. Wu et al. [71] proposed that microRNA-30 family members could inhibit the PAN-induced increase in Ca2+/calcineurin signaling in podocytes by targeting the inhibition of TRPC6 (Ca2+ channel), PPP3CA and PPP3CB (calcineurin α-catalytic subunit members [118]), PPP3R1 (calcineurin β-regulatory subunit member [118]), and NFATC3 (transcription factor). The antiproteinuric effect and podocyte protection of calcineurin inhibitors (CNI) have been confirmed [108]. Cyclosporine (CsA) and tacrolimus (FK506) are currently the most widely used calcineurin inhibitors [119]. CNI can ameliorate PAN-induced podocyte damage and apoptosis by inhibiting the CaN/NFAT/Angptl4 pathway [109]. Calcineurin can dephosphorylate synaptopodin, be degraded by CatL, and alter the morphology of the podocyte cytoskeleton. CNI destabilizes the podocyte cytoskeleton by targeting calcineurin [120]. In addition, one study showed that FK506 could reduce TRPC6 expression of the podocyte in DKD [110]. The specific mechanism for the beneficial effect of CNI on podocyte injury requires further exploration.

6. Conclusions

Ca2+ in podocytes can be stimulated by different signals and transmitted to multiple downstream signaling pathways. Podocytes sense intracellular Ca2+ concentrations through the CaSR, generating responses that regulate the balance of Ca2+, which is essential for the function of the glomerular filtration barrier. Similarly, the downstream signaling pathways mediated by the activity of Ca2+ channels are involved in various biological functions, such as the morphogenesis, structure, and survival of podocytes. In addition to the TRPCs widely studied in podocytes, recent studies have also suggested the importance of N-type calcium channels for Ca2+ transport in podocytes. RyR2, VDAC1, and SOCs are also important for maintaining organelle Ca2+ homeostasis. Increased Ca2+ influx and Ca2+ signaling activation in various podocyte-related diseases can achieve cell communications. FSGS, as one of the diseases most commonly manifested as SRNS, is associated with mutations in various Ca2+-signaling-pathway-related genes. Therapies targeting Ca2+ signaling have broad application prospects in patients with FSGS and also provide a novel option for therapy in patients with SRNS. Podocytopathy in DKD is widely recognized, and some progress has been made in the study of Ca2+ signaling in podocyte damage that has been induced by high levels of glucose and AngII. However, the effect of focusing Ca2+-signaling therapy on podocyte damage in DKD still needs further exploration. Abnormal IgG-antibody-mediated podocyte injury in LN and TG also gradually attracted the interest of researchers to the role of Ca2+; however, it too requires additional exploration. Ca2+ signaling targets may also be used as treatment strategies for immune-related kidney injuries. Currently, the therapeutic strategies targeting the Ca2+-signaling pathway focus on channel proteins, yet with the research progress on CaSRs and downstream signaling of Ca2+, they will soon be researched as potential therapeutic targets.

Author Contributions

YCT and LJY selected the topic. YCT wrote the paper. LJY provided guidance to the paper. LJY, YCT, HPS, LLS, QQL, LF, and MR participated in the discussion, revision, and correction of the manuscript. All authors contributed to editorial changes in the manuscript. All authors read and approved the final manuscript. All authors have participated sufficiently in the work to take public responsibility for appropriate portions of the content and agreed to be accountable for all aspects of the work in ensuring that questions related to its accuracy or integrity.

Ethics Approval and Consent to Participate

Not applicable.

Acknowledgment

Not applicable.

Funding

This research was supported by the National Natural Science Foundation of China (No.81974102).

Conflict of Interest

The authors declare no conflict of interest.

References
[1]
Sever S, Schiffer M. Actin dynamics at focal adhesions: a common endpoint and putative therapeutic target for proteinuric kidney diseases. Kidney International. 2018; 93: 1298–1307.
[2]
Schell C, Huber TB. The Evolving Complexity of the Podocyte Cytoskeleton. Journal of the American Society of Nephrology: JASN. 2017; 28: 3166–3174.
[3]
Nagata M. Podocyte injury and its consequences. Kidney International. 2016; 89: 1221–1230.
[4]
Kopp JB, Anders HJ, Susztak K, Podestà MA, Remuzzi G, Hildebrandt F, et al. Podocytopathies. Nature Reviews. Disease Primers. 2020; 6: 68.
[5]
Staruschenko A, Ma R, Palygin O, Dryer SE. Ion channels and channelopathies in glomeruli. Physiological Reviews. 2023; 103: 787–854.
[6]
Hannan FM, Kallay E, Chang W, Brandi ML, Thakker RV. The calcium-sensing receptor in physiology and in calcitropic and noncalcitropic diseases. Nature Reviews. Endocrinology. 2018; 15: 33–51.
[7]
Hall G, Wang L, Spurney RF. TRPC Channels in Proteinuric Kidney Diseases. Cells. 2019; 9: 44.
[8]
Djenoune L, Tomar R, Dorison A, Ghobrial I, Schenk H, Hegermann J, et al. Autonomous Calcium Signaling in Human and Zebrafish Podocytes Controls Kidney Filtration Barrier Morphogenesis. Journal of the American Society of Nephrology: JASN. 2021; 32: 1697–1712.
[9]
Bhargava R, Maeda K, Tsokos MG, Pavlakis M, Stillman IE, Tsokos GC. N-glycosylated IgG in patients with kidney transplants increases calcium/calmodulin kinase IV in podocytes and causes injury. American Journal of Transplantation: Official Journal of the American Society of Transplantation and the American Society of Transplant Surgeons. 2021; 21: 148–160.
[10]
Lu XY, Liu BC, Cao YZ, Song C, Su H, Chen G, et al. High glucose reduces expression of podocin in cultured human podocytes by stimulating TRPC6. American Journal of Physiology. Renal Physiology. 2019; 317: F1605–F1611.
[11]
Ma S, Xu J, Zheng Y, Li Y, Wang Y, Li H, et al. Qian Yang Yu Yin granule improves hypertensive renal damage: A potential role for TRPC6-CaMKKβ-AMPK-mTOR-mediated autophagy. Journal of Ethnopharmacology. 2023; 302: 115878.
[12]
Burford JL, Villanueva K, Lam L, Riquier-Brison A, Hackl MJ, Pippin J, et al. Intravital imaging of podocyte calcium in glomerular injury and disease. The Journal of Clinical Investigation. 2014; 124: 2050–2058.
[13]
Obeidová L, Reiterová J, Lněnička P, Štekrová J, Šafránková H, Kohoutová M, et al. TRPC6 gene variants in Czech adult patients with focal segmental glomerulosclerosis and minimal change disease. Folia Biologica. 2012; 58: 173–176.
[14]
Ilatovskaya DV, Levchenko V, Lowing A, Shuyskiy LS, Palygin O, Staruschenko A. Podocyte injury in diabetic nephropathy: implications of angiotensin II-dependent activation of TRPC channels. Scientific Reports. 2015; 5: 17637.
[15]
Piecha G, Kokeny G, Nakagawa K, Koleganova N, Geldyyev A, Berger I, et al. Calcimimetic R-568 or calcitriol: equally beneficial on progression of renal damage in subtotally nephrectomized rats. American Journal of Physiology. Renal Physiology. 2008; 294: F748–57.
[16]
Mühlig AK, Steingröver J, Heidelbach HS, Wingerath M, Sachs W, Hermans-Borgmeyer I, et al. The calcium-sensing receptor stabilizes podocyte function in proteinuric humans and mice. Kidney International. 2022; 101: 1186–1199.
[17]
Leach K, Hannan FM, Josephs TM, Keller AN, Møller TC, Ward DT, et al. International Union of Basic and Clinical Pharmacology. CVIII. Calcium-Sensing Receptor Nomenclature, Pharmacology, and Function. Pharmacological Reviews. 2020; 72: 558–604.
[18]
Chavez-Abiega S, Mos I, Centeno PP, Elajnaf T, Schlattl W, Ward DT, et al. Sensing Extracellular Calcium - An Insight into the Structure and Function of the Calcium-Sensing Receptor (CaSR). Advances in Experimental Medicine and Biology. 2020; 1131: 1031–1063.
[19]
Zhang L, Ji T, Wang Q, Meng K, Zhang R, Yang H, et al. Calcium-Sensing Receptor Stimulation in Cultured Glomerular Podocytes Induces TRPC6-Dependent Calcium Entry and RhoA Activation. Cellular Physiology and Biochemistry: International Journal of Experimental Cellular Physiology, Biochemistry, and Pharmacology. 2017; 43: 1777–1789.
[20]
Goel M, Sinkins WG, Zuo CD, Estacion M, Schilling WP. Identification and localization of TRPC channels in the rat kidney. American Journal of Physiology. Renal Physiology. 2006; 290: F1241–F1252.
[21]
Ilatovskaya DV, Levchenko V, Ryan RP, Cowley AW, Jr, Staruschenko A. NSAIDs acutely inhibit TRPC channels in freshly isolated rat glomeruli. Biochemical and Biophysical Research Communications. 2011; 408: 242–247.
[22]
Tian D, Jacobo SMP, Billing D, Rozkalne A, Gage SD, Anagnostou T, et al. Antagonistic regulation of actin dynamics and cell motility by TRPC5 and TRPC6 channels. Science Signaling. 2010; 3: ra77.
[23]
Ilatovskaya DV, Staruschenko A. TRPC6 channel as an emerging determinant of the podocyte injury susceptibility in kidney diseases. American Journal of Physiology. Renal Physiology. 2015; 309: F393–F397.
[24]
Gees M, Colsoul B, Nilius B. The role of transient receptor potential cation channels in Ca2+ signaling. Cold Spring Harbor Perspectives in Biology. 2010; 2: a003962.
[25]
Roos J, DiGregorio PJ, Yeromin AV, Ohlsen K, Lioudyno M, Zhang S, et al. STIM1, an essential and conserved component of store-operated Ca2+ channel function. The Journal of Cell Biology. 2005; 169: 435–445.
[26]
Li Y, Yang X, Shen Y. Structural Insights into Ca2+ Permeation through Orai Channels. Cells. 2021; 10: 3062.
[27]
Wang Y, Deng X, Gill DL. Calcium signaling by STIM and Orai: intimate coupling details revealed. Science Signaling. 2010; 3: pe42.
[28]
Park SJ, Kim Y, Yang SM, Henderson MJ, Yang W, Lindahl M, et al. Discovery of endoplasmic reticulum calcium stabilizers to rescue ER-stressed podocytes in nephrotic syndrome. Proceedings of the National Academy of Sciences of the United States of America. 2019; 116: 14154–14163.
[29]
Gong D, Chen X, Middleditch M, Huang L, Vazhoor Amarsingh G, Reddy S, et al. Quantitative proteomic profiling identifies new renal targets of copper(II)-selective chelation in the reversal of diabetic nephropathy in rats. Proteomics. 2009; 9: 4309–4320.
[30]
Bai L, Sun S, Sun Y, Wang F, Nishiyama A. N-type calcium channel and renal injury. International Urology and Nephrology. 2022; 54: 2871–2879.
[31]
Huang H-T, Lin X, Guo P-W, Pang J, Ma J, He L-L, et al. Expression and role of the TRPC family in TGF-β1-induced calcium influx in podocytes. Sheng Li Xue Bao. 2022; 74: 1005–1013. (In Chinese)
[32]
Polat OK, Isaeva E, Sudhini YR, Knott B, Zhu K, Noben M, et al. The small GTPase regulatory protein Rac1 drives podocyte injury independent of cationic channel protein TRPC5. Kidney International. 2023; 103: 1056–1062.
[33]
Persson PB. TRPs revisited. Acta Physiologica (Oxford, England). 2015; 214: 6–7.
[34]
Wang L, Jirka G, Rosenberg PB, Buckley AF, Gomez JA, Fields TA, et al. Gq signaling causes glomerular injury by activating TRPC6. The Journal of Clinical Investigation. 2015; 125: 1913–1926.
[35]
Ilatovskaya DV, Palygin O, Chubinskiy-Nadezhdin V, Negulyaev YA, Ma R, Birnbaumer L, et al. Angiotensin II has acute effects on TRPC6 channels in podocytes of freshly isolated glomeruli. Kidney International. 2014; 86: 506–514.
[36]
Forst AL, Olteanu VS, Mollet G, Wlodkowski T, Schaefer F, Dietrich A, et al. Podocyte Purinergic P2X4 Channels Are Mechanotransducers That Mediate Cytoskeletal Disorganization. Journal of the American Society of Nephrology: JASN. 2016; 27: 848–862.
[37]
Miao L, Wei D, Zhang Y, Liu J, Lu S, Zhang A, et al. Effects of stromal interaction molecule 1 or Orai1 overexpression on the associated proteins and permeability of podocytes. Nephrology (Carlton, Vic.). 2016; 21: 959–967.
[38]
Yan T, Zhao Y. Acetaldehyde induces phosphorylation of dynamin-related protein 1 and mitochondrial dysfunction via elevating intracellular ROS and Ca2+ levels. Redox Biology. 2020; 28: 101381.
[39]
Polat OK, Uno M, Maruyama T, Tran HN, Imamura K, Wong CF, et al. Contribution of Coiled-Coil Assembly to Ca2+/Calmodulin-Dependent Inactivation of TRPC6 Channel and its Impacts on FSGS-Associated Phenotypes. Journal of the American Society of Nephrology: JASN. 2019; 30: 1587–1603.
[40]
Greka A, Mundel P. Balancing calcium signals through TRPC5 and TRPC6 in podocytes. Journal of the American Society of Nephrology: JASN. 2011; 22: 1969–1980.
[41]
Hinkes B, Wiggins RC, Gbadegesin R, Vlangos CN, Seelow D, Nürnberg G, et al. Positional cloning uncovers mutations in PLCE1 responsible for a nephrotic syndrome variant that may be reversible. Nature Genetics. 2006; 38: 1397–1405.
[42]
Liu Z, Yang J, Zhang X, Xu P, Zhang T, Yang Z. Developmental changes in the expression and function of TRPC6 channels related the F-actin organization during differentiation in podocytes. Cell Calcium. 2015; 58: 541–548.
[43]
Ye Q, Lan B, Liu H, Persson PB, Lai EY, Mao J. A critical role of the podocyte cytoskeleton in the pathogenesis of glomerular proteinuria and autoimmune podocytopathies. Acta Physiologica (Oxford, England). 2022; 235: e13850.
[44]
Mulukala SKN, Kambhampati V, Qadri AH, Pasupulati AK. Evolutionary conservation of intrinsically unstructured regions in slit-diaphragm proteins. PLoS ONE. 2021; 16: e0254917.
[45]
Mukherjee R, Das A, Chakrabarti S, Chakrabarti O. Calcium dependent regulation of protein ubiquitination - Interplay between E3 ligases and calcium binding proteins. Biochimica et Biophysica Acta. Molecular Cell Research. 2017; 1864: 1227–1235.
[46]
Bhargava R, Lehoux S, Maeda K, Tsokos MG, Krishfield S, Ellezian L, et al. Aberrantly glycosylated IgG elicits pathogenic signaling in podocytes and signifies lupus nephritis. JCI Insight. 2021; 6: e147789.
[47]
Hinkovska-Galcheva V, Clark A, VanWay S, Huang JB, Hiraoka M, Abe A, et al. Ceramide kinase promotes Ca2+ signaling near IgG-opsonized targets and enhances phagolysosomal fusion in COS-1 cells. Journal of Lipid Research. 2008; 49: 531–542.
[48]
Qu L, Li Y, Pan X, Zhang P, LaMotte RH, Ma C. Transient receptor potential canonical 3 (TRPC3) is required for IgG immune complex-induced excitation of the rat dorsal root ganglion neurons. The Journal of Neuroscience: the Official Journal of the Society for Neuroscience. 2012; 32: 9554–9562.
[49]
Yu SMW, Nissaisorakarn P, Husain I, Jim B. Proteinuric Kidney Diseases: A Podocyte’s Slit Diaphragm and Cytoskeleton Approach. Frontiers in Medicine. 2018; 5: 221.
[50]
Maeda K, Otomo K, Yoshida N, Abu-Asab MS, Ichinose K, Nishino T, et al. CaMK4 compromises podocyte function in autoimmune and nonautoimmune kidney disease. The Journal of Clinical Investigation. 2018; 128: 3445–3459.
[51]
Schaldecker T, Kim S, Tarabanis C, Tian D, Hakroush S, Castonguay P, et al. Inhibition of the TRPC5 ion channel protects the kidney filter. The Journal of Clinical Investigation. 2013; 123: 5298–5309.
[52]
Asano-Matsuda K, Ibrahim S, Takano T, Matsuda J. Role of Rho GTPase Interacting Proteins in Subcellular Compartments of Podocytes. International Journal of Molecular Sciences. 2021; 22: 3656.
[53]
Kim JH, Hwang KH, Dang BTN, Eom M, Kong ID, Gwack Y, et al. Insulin-activated store-operated Ca2+ entry via Orai1 induces podocyte actin remodeling and causes proteinuria. Nature Communications. 2021; 12: 6537.
[54]
Verheijden KAT, Sonneveld R, Bakker-van Bebber M, Wetzels JFM, van der Vlag J, Nijenhuis T. The Calcium-Dependent Protease Calpain-1 Links TRPC6 Activity to Podocyte Injury. Journal of the American Society of Nephrology: JASN. 2018; 29: 2099–2109.
[55]
Farmer LK, Rollason R, Whitcomb DJ, Ni L, Goodliff A, Lay AC, et al. TRPC6 Binds to and Activates Calpain, Independent of Its Channel Activity, and Regulates Podocyte Cytoskeleton, Cell Adhesion, and Motility. Journal of the American Society of Nephrology: JASN. 2019; 30: 1910–1924.
[56]
Faul C, Asanuma K, Yanagida-Asanuma E, Kim K, Mundel P. Actin up: regulation of podocyte structure and function by components of the actin cytoskeleton. Trends in Cell Biology. 2007; 17: 428–437.
[57]
Oh J, Beckmann J, Bloch J, Hettgen V, Mueller J, Li L, et al. Stimulation of the calcium-sensing receptor stabilizes the podocyte cytoskeleton, improves cell survival, and reduces toxin-induced glomerulosclerosis. Kidney International. 2011; 80: 483–492.
[58]
Tossidou I, Teng B, Worthmann K, Müller-Deile J, Jobst-Schwan T, Kardinal C, et al. Tyrosine Phosphorylation of CD2AP Affects Stability of the Slit Diaphragm Complex. Journal of the American Society of Nephrology: JASN. 2019; 30: 1220–1237.
[59]
Ruby M, Gifford CC, Pandey R, Raj VS, Sabbisetti VS, Ajay AK. Autophagy as a Therapeutic Target for Chronic Kidney Disease and the Roles of TGF-β1 in Autophagy and Kidney Fibrosis. Cells. 2023; 12: 412.
[60]
Bao H, Ge Y, Peng A, Gong R. Fine-tuning of NFκB by glycogen synthase kinase 3β directs the fate of glomerular podocytes upon injury. Kidney International. 2015; 87: 1176–1190.
[61]
Bensaada I, Robin B, Perez J, Salemkour Y, Chipont A, Camus M, et al. Calpastatin prevents Angiotensin II-mediated podocyte injury through maintenance of autophagy. Kidney International. 2021; 100: 90–106.
[62]
Gao N, Wang H, Yin H, Yang Z. Angiotensin II induces calcium-mediated autophagy in podocytes through enhancing reactive oxygen species levels. Chemico-biological Interactions. 2017; 277: 110–118.
[63]
Nijenhuis T, Sloan AJ, Hoenderop JGJ, Flesche J, van Goor H, Kistler AD, et al. Angiotensin II contributes to podocyte injury by increasing TRPC6 expression via an NFAT-mediated positive feedback signaling pathway. The American Journal of Pathology. 2011; 179: 1719–1732.
[64]
Xu H, Guan N, Ren YL, Wei QJ, Tao YH, Yang GS, et al. IP_3R-Grp75-VDAC1-MCU calcium regulation axis antagonists protect podocytes from apoptosis and decrease proteinuria in an Adriamycin nephropathy rat model. BMC Nephrology. 2018; 19: 140.
[65]
Xu S, Nam SM, Kim JH, Das R, Choi SK, Nguyen TT, et al. Palmitate induces ER calcium depletion and apoptosis in mouse podocytes subsequent to mitochondrial oxidative stress. Cell Death & Disease. 2015; 6: e1976.
[66]
Yoshimura Y, Nishinakamura R. Podocyte development, disease, and stem cell research. Kidney International. 2019; 96: 1077–1082.
[67]
Riehle M, Büscher AK, Gohlke BO, Kaßmann M, Kolatsi-Joannou M, Bräsen JH, et al. TRPC6 G757D Loss-of-Function Mutation Associates with FSGS. Journal of the American Society of Nephrology: JASN. 2016; 27: 2771–2783.
[68]
Liu Z, Zhang H, Zhao S, Zhang Q, Zhang R, Han Y, et al. Novel gain-of-function mutation of TRPC6 Q134P contributes to late onset focal segmental glomerulosclerosis in a Chinese pedigree. Nephrology (Carlton, Vic.). 2021; 26: 1018–1025.
[69]
Gigante M, Caridi G, Montemurno E, Soccio M, d’Apolito M, Cerullo G, et al. TRPC6 mutations in children with steroid-resistant nephrotic syndrome and atypical phenotype. Clinical Journal of the American Society of Nephrology: CJASN. 2011; 6: 1626–1634.
[70]
Hofstra JM, Lainez S, van Kuijk WHM, Schoots J, Baltissen MPA, Hoefsloot LH, et al. New TRPC6 gain-of-function mutation in a non-consanguineous Dutch family with late-onset focal segmental glomerulosclerosis. Nephrology, Dialysis, Transplantation: Official Publication of the European Dialysis and Transplant Association - European Renal Association. 2013; 28: 1830–1838.
[71]
Wu J, Zheng C, Wang X, Yun S, Zhao Y, Liu L, et al. MicroRNA-30 family members regulate calcium/calcineurin signaling in podocytes. The Journal of Clinical Investigation. 2015; 125: 4091–4106.
[72]
Chiluiza D, Krishna S, Schumacher VA, Schlöndorff J. Gain-of-function mutations in transient receptor potential C6 (TRPC6) activate extracellular signal-regulated kinases 1/2 (ERK1/2). The Journal of Biological Chemistry. 2013; 288: 18407–18420.
[73]
Wang Y, Jarad G, Tripathi P, Pan M, Cunningham J, Martin DR, et al. Activation of NFAT signaling in podocytes causes glomerulosclerosis. Journal of the American Society of Nephrology: JASN. 2010; 21: 1657–1666.
[74]
Franco SJ, Rodgers MA, Perrin BJ, Han J, Bennin DA, Critchley DR, et al. Calpain-mediated proteolysis of talin regulates adhesion dynamics. Nature Cell Biology. 2004; 6: 977–983.
[75]
Wang L, Chang J-H, Buckley AF, Spurney RF. Corrigendum to “Wang L, Chang J-H, Buckley AF, Spurney RF. Knockout of TRPC6 promotes insulin resistance and exacerbates glomerular injury in Akita mice.” Kidney Int. 2019;95:321-332. Kidney International. 2020; 97: 808–809.
[76]
Audzeyenka I, Bierżyńska A, Lay AC. Podocyte Bioenergetics in the Development of Diabetic Nephropathy: The Role of Mitochondria. Endocrinology. 2022; 163: bqab234.
[77]
Yu H, Chen Y, Ma H, Wang Z, Zhang R, Jiao J. TRPC6 mediates high glucose-induced mitochondrial fission through activation of CDK5 in cultured human podocytes. Frontiers in Physiology. 2022; 13: 984760.
[78]
Tao Y, Yazdizadeh Shotorbani P, Inman D, Das-Earl P, Ma R. Store-operated Ca2+ entry inhibition ameliorates high glucose and ANG II-induced podocyte apoptosis and mitochondrial damage. American Journal of Physiology. Renal Physiology. 2023; 324: F494–F504.
[79]
Rizzuto R, De Stefani D, Raffaello A, Mammucari C. Mitochondria as sensors and regulators of calcium signalling. Nature Reviews. Molecular Cell Biology. 2012; 13: 566–578.
[80]
Giacomello M, Pellegrini L. The coming of age of the mitochondria-ER contact: a matter of thickness. Cell Death and Differentiation. 2016; 23: 1417–1427.
[81]
Wei X, Wei X, Lu Z, Li L, Hu Y, Sun F, et al. Activation of TRPV1 channel antagonizes diabetic nephropathy through inhibiting endoplasmic reticulum-mitochondria contact in podocytes. Metabolism: Clinical and Experimental. 2020; 105: 154182.
[82]
Marchi S, Patergnani S, Missiroli S, Morciano G, Rimessi A, Wieckowski MR, et al. Mitochondrial and endoplasmic reticulum calcium homeostasis and cell death. Cell Calcium. 2018; 69: 62–72.
[83]
Yang D, Livingston MJ, Liu Z, Dong G, Zhang M, Chen JK, et al. Autophagy in diabetic kidney disease: regulation, pathological role and therapeutic potential. Cellular and Molecular Life Sciences: CMLS. 2018; 75: 669–688.
[84]
Shi W, Huang Y, Zhao X, Xie Z, Dong W, Li R, et al. Histone deacetylase 4 mediates high glucose-induced podocyte apoptosis via upregulation of calcineurin. Biochemical and Biophysical Research Communications. 2020; 533: 1061–1068.
[85]
Tao Y, Chaudhari S, Shotorbani PY, Ding Y, Chen Z, Kasetti R, et al. Enhanced Orai1-mediated store-operated Ca2+ channel/calpain signaling contributes to high glucose-induced podocyte injury. The Journal of Biological Chemistry. 2022; 298: 101990.
[86]
Ohno S, Yokoi H, Mori K, Kasahara M, Kuwahara K, Fujikura J, et al. Ablation of the N-type calcium channel ameliorates diabetic nephropathy with improved glycemic control and reduced blood pressure. Scientific Reports. 2016; 6: 27192.
[87]
Sakhi H, Moktefi A, Bouachi K, Audard V, Hénique C, Remy P, et al. Podocyte Injury in Lupus Nephritis. Journal of Clinical Medicine. 2019; 8: 1340.
[88]
Ichinose K, Ushigusa T, Nishino A, Nakashima Y, Suzuki T, Horai Y, et al. Lupus Nephritis IgG Induction of Calcium/Calmodulin-Dependent Protein Kinase IV Expression in Podocytes and Alteration of Their Function. Arthritis & Rheumatology (Hoboken, N.J.). 2016; 68: 944–952.
[89]
Yang Y, Hodgin JB, Afshinnia F, Wang SQ, Wickman L, Chowdhury M, et al. The two kidney to one kidney transition and transplant glomerulopathy: a podocyte perspective. Journal of the American Society of Nephrology: JASN. 2015; 26: 1450–1465.
[90]
Costantino VV, Gil Lorenzo AF, Bocanegra V, Vallés PG. Molecular Mechanisms of Hypertensive Nephropathy: Renoprotective Effect of Losartan through Hsp70. Cells. 2021; 10: 3146.
[91]
Lin H, Geurts F, Hassler L, Batlle D, Mirabito Colafella KM, Denton KM, et al. Kidney Angiotensin in Cardiovascular Disease: Formation and Drug Targeting. Pharmacological Reviews. 2022; 74: 462–505.
[92]
Lei B, Nakano D, Fujisawa Y, Liu Y, Hitomi H, Kobori H, et al. N-type calcium channel inhibition with cilnidipine elicits glomerular podocyte protection independent of sympathetic nerve inhibition. Journal of Pharmacological Sciences. 2012; 119: 359–367.
[93]
Fan YY, Kohno M, Nakano D, Ohsaki H, Kobori H, Suwarni D, et al. Cilnidipine suppresses podocyte injury and proteinuria in metabolic syndrome rats: possible involvement of N-type calcium channel in podocyte. Journal of Hypertension. 2010; 28: 1034–1043.
[94]
Golosova D, Palygin O, Bohovyk R, Klemens CA, Levchenko V, Spires DR, et al. Role of opioid signaling in kidney damage during the development of salt-induced hypertension. Life Science Alliance. 2020; 3: e202000853.
[95]
Chen L, Tang YL, Liu ZH, Pan Y, Jiao RQ, Kong LD. Atractylodin inhibits fructose-induced human podocyte hypermotility via anti-oxidant to down-regulate TRPC6/p-CaMK4 signaling. European Journal of Pharmacology. 2021; 913: 174616.
[96]
Chen L, Yang J, Zhao SJ, Li TS, Jiao RQ, Kong LD. Atractylodis rhizoma water extract attenuates fructose-induced glomerular injury in rats through anti-oxidation to inhibit TRPC6/p-CaMK4 signaling. Phytomedicine: International Journal of Phytotherapy and Phytopharmacology. 2021; 91: 153643.
[97]
Zang Y, Liu S, Cao A, Shan X, Deng W, Li Z, et al. Astragaloside IV inhibits palmitic acid-induced apoptosis through regulation of calcium homeostasis in mice podocytes. Molecular Biology Reports. 2021; 48: 1453–1464.
[98]
Liu X, Zhou QG, Zhu XC, Xie L, Cai BC. Screening for Potential Active Components of Fangji Huangqi Tang on the Treatment of Nephrotic Syndrome by Using Integrated Metabolomics Based on “Correlations Between Chemical and Metabolic Profiles”. Frontiers in Pharmacology. 2019; 10: 1261.
[99]
Yu J, Zhu C, Yin J, Yu D, Wan F, Tang X, et al. Tetrandrine Suppresses Transient Receptor Potential Cation Channel Protein 6 Overexpression- Induced Podocyte Damage via Blockage of RhoA/ROCK1 Signaling. Drug Design, Development and Therapy. 2020; 14: 361–370.
[100]
Ding Y, Tang X, Wang Y, Yu D, Zhu C, Yu J. Tetrandrine alleviates podocyte injury via calcium-dependent calpain-1 signaling blockade. BMC Complementary Medicine and Therapies. 2021; 21: 296.
[101]
Zhang R, Wang X, Gao Q, Jiang H, Zhang S, Lu M, et al. Taurine Supplementation Reverses Diabetes-Induced Podocytes Injury via Modulation of the CSE/TRPC6 Axis and Improvement of Mitochondrial Function. Nephron. 2020; 144: 84–95.
[102]
Urban N, Neuser S, Hentschel A, Köhling S, Rademann J, Schaefer M. Pharmacological inhibition of focal segmental glomerulosclerosis-related, gain of function mutants of TRPC6 channels by semi-synthetic derivatives of larixol. British Journal of Pharmacology. 2017; 174: 4099–4122.
[103]
Szrejder M, Rachubik P, Rogacka D, Audzeyenka I, Rychłowski M, Kreft E, et al. Metformin reduces TRPC6 expression through AMPK activation and modulates cytoskeleton dynamics in podocytes under diabetic conditions. Biochimica et Biophysica Acta. Molecular Basis of Disease. 2020; 1866: 165610.
[104]
Kolligundla LP, Kavvuri R, Singh AK, Mukhi D, Pasupulati AK. Metformin prevents hypoxia-induced podocyte injury by regulating the ZEB2/TG2 axis. Nephrology (Carlton, Vic.). 2023; 28: 60–71.
[105]
Youssef N, Noureldein M, Njeim R, Ghadieh HE, Harb F, Azar ST, et al. Reno-Protective Effect of GLP-1 Receptor Agonists in Type1 Diabetes: Dual Action on TRPC6 and NADPH Oxidases. Biomedicines. 2021; 9: 1360.
[106]
Sonneveld R, Hoenderop JG, Isidori AM, Henique C, Dijkman HB, Berden JH, et al. Sildenafil Prevents Podocyte Injury via PPAR-γ-Mediated TRPC6 Inhibition. Journal of the American Society of Nephrology: JASN. 2017; 28: 1491–1505.
[107]
‘t Hart D, Li J, van der Vlag J, Nijenhuis T. Repurposing Riociguat to Target a Novel Paracrine Nitric Oxide-TRPC6 Pathway to Prevent Podocyte Injury. International Journal of Molecular Sciences. 2021; 22: 12485.
[108]
Malakasioti G, Iancu D, Milovanova A, Tsygin A, Horinouchi T, Nagano C, et al. A multicenter retrospective study of calcineurin inhibitors in nephrotic syndrome secondary to podocyte gene variants. Kidney International. 2023; 103: 962–972.
[109]
Shen X, Zhang Y, Lin C, Weng C, Wang Y, Feng S, et al. Calcineurin inhibitors ameliorate PAN-induced podocyte injury through the NFAT-Angptl4 pathway. The Journal of Pathology. 2020; 252: 227–238.
[110]
Ma R, Wang Y, Xu Y, Wang R, Wang X, Yu N, et al. Tacrolimus Protects Podocytes from Apoptosis via Downregulation of TRPC6 in Diabetic Nephropathy. Journal of Diabetes Research. 2021; 2021: 8832114.
[111]
Huang L, Zhao YJ, Dong QR, Hu GC. A study of altered calcium sensing system caused primary membranous nephropathy to end-stage renal failure. Biomedicine & Pharmacotherapy. 2021; 141: 111777.
[112]
Yadav S, Gupta K, Deshmukh K, Bhardwaj L, Dahiya A, Krishan P, et al. Calcium sensing receptor as a novel target for treatment of sepsis induced cardio-renal syndrome: Need for exploring mechanisms. Drug Development Research. 2021; 82: 305–308.
[113]
Wang D, Wang Q, Ji T, Yang H, Kong F, Jiao J. The role of TRPC6 in α1-AR activation-induced calcium signal changes in human podocytes. Annals of Palliative Medicine. 2020; 9: 1596–1605.
[114]
Britto-Júnior J, Ribeiro A, Ximenes L, Lima AT, Jacintho FF, Fregonesi A, et al. Alpha1-adrenergic antagonists block 6-nitrodopamine contractions on the rat isolated epididymal vas deferens. European Journal of Pharmacology. 2022; 915: 174716.
[115]
Tajiri K, Guichard JB, Qi X, Xiong F, Naud P, Tardif JC, et al. An N-/L-type calcium channel blocker, cilnidipine, suppresses autonomic, electrical, and structural remodelling associated with atrial fibrillation. Cardiovascular Research. 2019; 115: 1975–1985.
[116]
Fujita T, Ando K, Nishimura H, Ideura T, Yasuda G, Isshiki M, et al. Antiproteinuric effect of the calcium channel blocker cilnidipine added to renin-angiotensin inhibition in hypertensive patients with chronic renal disease. Kidney International. 2007; 72: 1543–1549.
[117]
Morimoto S, Yano Y, Maki K, Iwasaka T. Renal and vascular protective effects of cilnidipine in patients with essential hypertension. Journal of Hypertension. 2007; 25: 2178–2183.
[118]
Roy J, Cyert MS. Identifying New Substrates and Functions for an Old Enzyme: Calcineurin. Cold Spring Harbor Perspectives in Biology. 2020; 12: a035436.
[119]
Zhou JJ, Shao JY, Chen SR, Pan HL. Calcineurin Controls Hypothalamic NMDA Receptor Activity and Sympathetic Outflow. Circulation Research. 2022; 131: 345–360.
[120]
Vassiliadis J, Bracken C, Matthews D, O’Brien S, Schiavi S, Wawersik S. Calcium mediates glomerular filtration through calcineurin and mTORC2/Akt signaling. Journal of the American Society of Nephrology: JASN. 2011; 22: 1453–1461.

Publisher’s Note: IMR Press stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share
Back to top