Next Article in Journal
Dual Immune Regulatory Roles of Interleukin-33 in Pathological Conditions
Next Article in Special Issue
The Role of the Epidermal Growth Factor Receptor in Diabetic Kidney Disease
Previous Article in Journal
The Power of Gene Technologies: 1001 Ways to Create a Cell Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Contribution of Lipotoxicity to Diabetic Kidney Disease

by
Jeffrey R. Schelling
Department of Physiology and Biophysics, School of Medicine, Case Western Reserve University, Cleveland, OH 44106-4921, USA
Submission received: 16 September 2022 / Revised: 2 October 2022 / Accepted: 12 October 2022 / Published: 14 October 2022

Abstract

:
Lipotoxicity is a fundamental pathophysiologic mechanism in diabetes and non-alcoholic fatty liver disease and is now increasingly recognized in diabetic kidney disease (DKD) pathogenesis. This review highlights lipotoxicity pathways in the podocyte and proximal tubule cell, which are arguably the two most critical sites in the nephron for DKD. The discussion focuses on membrane transporters and lipid droplets, which represent potential therapeutic targets, as well as current and developing pharmacologic approaches to reduce renal lipotoxicity.

1. Lipotoxicity Overview

Lipotoxicity is a metabolic condition that results from the intracellular accumulation of toxic lipid intermediates in non-adipose tissue, leading to cellular dysfunction and potentially cell death (lipoapoptosis) [1]. The prevalence of lipotoxicity is estimated to be 25% in adult populations [2]. The most common affected tissues are kidney, liver (leading to hepatosteatosis), heart (leading to congestive failure), pancreas (leading to β-cell death and diabetes), and skeletal muscle [3]. The parallels between fatty liver disease and diabetic kidney disease (DKD) are increasingly recognized [4]. In liver, there is a well-described spectrum from hepatosteatosis [lipid droplet (LD) accumulation, no inflammation or fibrosis) to steatohepatitis [hepatocyte ballooning, smaller LDs, increased toxic fatty acid (FA) metabolites, no fibrosis] to cirrhosis (fibrosis, fewer LDs) [5,6]. A similar scenario is likely to be apt with DKD, wherein LDs are associated with cytoprotection, and toxic FA metabolites lead to tubular atrophy, interstitial fibrosis and DKD progression.
In kidney, the association between renal tubular epithelial cell lipid deposition and DKD was described many years ago [7,8], and lipotoxicity has been implicated in chronic kidney disease (CKD) progression in humans and animal models [9,10,11,12,13,14]. Nevertheless, the impact of lipotoxicity in human DKD pathogenesis continues to be underestimated because clinical renal biopsies are typically not performed in diabetic people, and biopsy tissue is not routinely processed for lipid detection.
Diabetes is a lipogenic state and the toxic effects of excess lipids are buffered by uptake of FAs by adipocytes. However, when lipid storage capacity is exceeded, FAs are released to the circulation, and either metabolized by the liver and/or stored ectopically in other organs. Lipotoxicity has been associated primarily with ramifications of long-chain (C12-C22) FA metabolite accumulation, and cytotoxicity is proportional to acyl chain length and carbon bond saturation [15,16]. This is relevant to diabetic complications, since plasma concentrations of saturated FAs palmitate (C16:0) and stearate (C18:0) are increased in diabetes [17], and correlate with CKD stage [18]. Palmitate is particularly noxious, since it does not readily incorporate into triacylglycerol (TAG) or cytoprotective LDs, and is instead shunted to ceramide formation in the endoplasmic reticulum (ER) or stimulation of reactive oxygen species (ROS) generation [19], both of which lead to apoptosis [20,21,22,23]. Some, but not all studies indicate that unsaturated FA exposure may be cytoprotective [24,25], perhaps due to preferential partitioning to LDs [19,26,27] or through effects on proteasome function [28].
The mechanism by which resident kidney cells accumulate FAs and cholesterol is multifactorial, and include increased lipogenesis [29], increased uptake from extracellular compartments [30,31], and decreased efflux from intracellular to extracellular compartments [32]. β-oxidation may also be decreased [33], though this is an inconsistent finding [34]. Cell-specific mechanisms are discussed in subsequent sections devoted to podocytes and proximal tubule epithelial cells, the two most important cell types in DKD pathogenesis.
There are multiple intracellular pathways that can be activated to minimize lipotoxicity, most notably the sequestration of FAs and cholesterol in cytosolic LDs, which is discussed in detail in a separate section of this review. Unlike adipocytes, where the cytoplasm is almost entirely filled with LDs, epithelial cells have a limited capacity to buffer excess intracellular lipids through LD accrual. Therefore, once the threshold for FA metabolism and storage has been exceeded, the residual FAs exert lipotoxic effects. If these cytoprotective pathways are surmounted, the released lipid metabolites, including ceramides, sphingolipids, diacylglycerols and long-chain acyl-CoAs, can lead to membrane disruption, aberrant lipid modification and malfunction of proteins, such as transporters and signaling enzymes, ROS generation, ER stress, impaired autophagy and increased apoptosis. These downstream sequelae have recently been extensively reviewed [35,36]; this review instead focuses on upstream lipotoxicity triggers, which may be amenable to interventions that abrogate DKD progression.

2. Lipotoxicity of Podocytes

Podocytes are highly specialized and terminally differentiated (incapable of self-regeneration) cells, which are critical for normal glomerular function. The loss of podocytes leads to reduced glomerular basement membrane coverage, which manifests initially as albuminuria, and ultimately glomerulosclerosis. In DKD and non-DKD models the accumulation of both cholesterol ester and FA metabolites in podocytes has been implicated in the pathogenesis of glomerular dysfunction (Figure 1).

2.1. Podocyte Lipotoxicity Due to Altered Cholesterol Metabolism

Podocytes are unique by virtue of plasma membrane projections that form the slit diaphragm. These membrane domains are enriched with cholesterol-concentrated lipid rafts, which aggregate proteins that regulate the glomerular filtration barrier, such as nephrin, podocin, CD2-associated protein, α-actinin-4 and TRPC6 [37,38]. Multiple studies have demonstrated alterations in plasma membrane lipid components in DKD. Microarray studies from human biopsies identified significant upregulation of mRNA expression of acid sphingomyelinase-like phosphodiesterase 3b (SMPDL3b), which localizes to lipid rafts, and catalyzes sphingomyelin catabolism [39]. Biochemical ramifications included decreased ceramide-1-phosphate and perhaps increased free ceramide [40], which induces apoptosis (Figure 1). Genetic deletion or chemical inhibition of SMPDL3b inhibited podocyte apoptosis and albuminuria in Leprdb/db mice [39], which phenocopy type 2 diabetes and some aspects of DKD. A subsequent report linked increased SMPDL3b expression to altered caveolin-dependent insulin receptor function [40]. Podocyte-specific deletion of SMPDL3b or supplementation of the downstream product C16:0 ceramide-1-phosphate also rescued podocyte insulin signaling, apoptosis and albuminuria [40]. Interestingly, opposite effects were observed with adriamycin-induced podocyte injury, suggesting a SMPDL3b mechanism may not be common to all podocytopathies, perhaps due to differences in lipid concentrations and exposure.
Two groups have shown that cholesterol accumulates in the cytoplasm of podocytes in humans and animal models of DKD [41,42,43] (Figure 1). Gene expression and functional experiments support both enhanced cholesterol uptake, as well as decreased efflux from decreased ATP-binding cassette A1 (ABCA1) cholesterol transporter activity [41,42,43]. The mechanism by which cholesterol-containing lipoproteins would achieve access to podocyte LDL receptors in vivo remains unclear. A series of studies demonstrated decreased podocyte expression of ABCA1 in mouse (BTBRob/ob, BKSdb/db and streptozotocin-treated) models of diabetes [41,42,43]. Increased podocyte ABCA1 expression in transgenic or pharmacologically stimulated mice diminished albuminuria [42,43]. Furthermore, induction of cholesterol efflux with cyclodextrin improved albuminuria, as well as glycemic control, while treatment with HMG-CoA reductase inhibitors had no effect [32], suggesting that decreased cholesterol efflux, rather than enhanced cholesterol synthesis, was the mechanism of cholesterol accumulation. However, recent studies indicate that the ABCA1-dependent mechanism may be due to accumulation of cardiolipin, rather than cholesterol [43]. While these studies collectively support ABCA1-associated podocyte lipotoxicity in DKD pathophysiology, confirmation in DKD animal models that exhibit GFR decline or tubulointerstitial disease [14,44,45], which is tightly associated with GFR decline, is warranted.

2.2. Podocyte Lipotoxicity Due to Altered Fatty Acid Metabolism

Podocyte FAs may accumulate due to enhanced uptake, increased synthesis, or decreased degradation [46] (Figure 1). In studies from Fu et al., expression of junctional adhesion molecule-like protein (JAML) was noted to be increased in glomeruli from diabetic patients and two diabetic mouse models, and inversely correlated with GFR [47]. Importantly, podocyte-specific JAML deletion ameliorated lipid accumulation, glomerular histopathology and albuminuria by inhibiting sirtuin-1-dependent, sterol regulatory element-binding protein (SREBP)1-mediated FA synthesis [47].
Exposure of podocytes to the saturated, long-chain saturated FA palmitate is toxic, though compensatory upregulation of desaturase enzymes may initially limit toxicity in DKD [48]. Chung et al. [30] demonstrated that medium and long-chain FA uptake via the G-protein-coupled FFA1 receptor mediated cytoskeletal rearrangements that mimic podocyte foot process effacement (Figure 1). FFA1-dependent FA uptake may also precipitate albuminuria through an indirect mechanism, by inducing angiopoietin-like 4 [49]. FA uptake in podocytes is additionally mediated by the CD36 scavenger receptor; palmitate exposure to cultured podocytes stimulated CD36 expression, as well as ROS-dependent apoptosis [50]. Podocyte FA accumulation was also noted in diabetic BKSdb/db mice, due to a mechanism involving paracrine VEGF-B-mediated increase in podocyte FATP4 expression [51].

2.3. Lipotoxicity Involving the Tubulointerstitium

Proximal tubule lipid accumulation in DKD was first described by Kimmelstiel [7], and a link between tubule lipotoxicity and DKD is well established in humans [41], as well as rat [52] and mouse [29,33,53,54,55,56] models of type 1 and 2 diabetes. The prevailing model of proximal tubule lipotoxicity is predicated upon leakage of albumin-bound FAs across the injured podocyte and glomerular basement membrane, with subsequent uptake by apical proximal tubule transporters (Figure 2).
Apical proximal tubule glucose reabsorption is mediated by SGLT and GLUT transporters, but the purpose is for glucose reclamation, rather than metabolism. Although capable of modulating to glucose utilization, the proximal tubule is usually gluconeogenic. However, proximal tubules have the capacity to utilize glucose and gluconeogenic substrates for ATP generation [57,58,59,60,61,62,63], and one purported mechanism of SGLT2 inhibitor efficacy is through inhibition of the proximal tubule shift to glucose utilization and glycolysis [61,64]. Under normal circumstances metabolically demanding proximal tubules rely upon FAs, which are more efficient than glucose [65], as the main substrate for ATP generation [66,67]. Circulating FAs are transported across the basolateral membrane by undefined mechanisms (Figure 2). Following uptake, FAs are rapidly converted to acyl-CoA molecules by acyl-CoA synthetases, then preferentially shuttled to the outer mitochondrial membrane, and catalyzed by carnitine palmitoyl transferase-1 (CPT1), the rate-limiting step in FA metabolism. This conversion to acylcarnitine permits FA to traverse the mitochondrial matrix and inner mitochondrial membrane. Conversion of acylcarnitine back to long-chain acyl-CoA by CPT2, and then long-chain FAs, permits FAs to then undergo β-oxidation, ultimately yielding acetyl CoA plus FADH2 and NADH. This is then followed by the synthesis of citrate from acetyl CoA plus oxaloacetate, which can then enter the Krebs cycle, and leads to ATP generation.
Once proximal tubule cell metabolic needs are met, unless the FAs are sequestered, or metabolized without significant ROS generation, excess FAs and toxic metabolites lead to tubular atrophy through mitochondrial and ER stress and lipoapoptosis. The mechanism by which FAs accumulate in the proximal tubule is multifactorial, with evidence for increased uptake [31,41,63,68,69] and increased synthesis, by SREBPs, fatty acid synthase, carbohydrate response element-binding protein (ChREBP), peroxisome proliferator-activated receptor-α (PPARα), and farnesoid X receptor (FXR) deficiency [29,52,53,54,55,56]. These pathways are not mutually exclusive, but likely additive, and the consequence is the accumulation of toxic long-chain FA and metabolites, such as long-chain acyl-CoAs (LC-CoA), ceramides and diacylglycerol (DAG) [13,63,70] (Figure 2). Some [29,33,41], but not all [34,63,71] studies have also shown decreased β-oxidation, which is reflected in metabolomic data from diabetic sera [18].
FAs circulate either esterified to a glycerol backbone as triacylglycerol (TAG), or in a non-esterified form bound to albumin. Under normal circumstances TAG and non-esterified FA are too large to cross the glomerular filtration barrier. However, in “nephrotic” glomerular diseases, of which DKD is a prototype, FA-bound albumin traverses the injured GBM. Therefore, an important consequence of DKD is the aberrant filtration and exposure of albumin-bound long-chain FAs to the apical surface of the proximal tubule (Figure 2). Dextran bead clearance experiments demonstrate that molecules with an effective radius equivalent to albumin (3.5 nm) are permeable to glomeruli with DKD [72,73], whereas TAG within lipoproteins (>11 nm) would still be too large. Although DKD may progress without albuminuria [74], it remains one of the biggest risks for progression to end stage kidney disease [75]. Albumin per se is not cytotoxic to the proximal tubule [13,68,69,76], but FA bound to albumin instigate apoptosis in the proximal tubule [9,10,11] (Figure 2), as well as in other tissues [3,10,34,77].
Approximately 50% of apical proximal tubule NEFA uptake is mediated by the high affinity, low capacity transporter, FATP2 [31]. Under normal circumstances FATP2 is a scavenger for low concentrations of filtered FAs. However, in pathologic states, in which the glomerular filtration barrier is injured, permitting albumin-bound FA filtration and exposure to the AP proximal tubule membrane, consequent FATP2-dependent FA uptake leads to proximal tubular epithelial cell apoptosis and tubular atrophy [31]. As previously mentioned, proximal tubules utilize FAs as the major substrate for ATP generation, and multiple groups have determined that β-oxidation is diminished in DKD [29,33,41], resulting in decreased ATP generation in this energy-demanding nephron segment. Furthermore, when unmetabolized FAs are combined with enhanced FATP2-dependent FA uptake, intracellular FAs and FA metabolites, such as long-chain acyl-CoAs, accumulate and contribute to lipotoxicity [13].
Global FATP2 gene (Slc27a2) deletion combined with genetic or inducible mouse models that phenocopy progressive DKD, resulted in abrogation of tubular atrophy and restoration of GFR [14]. Although amelioration of DKD could be due to deletion of proximal tubule FATP2, the global FATP2 knockout mice also achieved dramatic reductions in plasma glucose, which may exert indirect kidney benefits [78]. Furthermore, FATP2 deletion does not completely eliminate apical proximal tubule FA uptake [31]. The residual FA transporters are incompletely characterized; FATP3 is also expressed along the apical proximal tubule membrane, though not nearly as abundantly as FATP2 (Khan et al., unpublished observations). In addition to mediating bulk FA uptake, FATP2 physically interacts with ceramide synthase 2, and contributes catalytically to increased synthesis of potentially cytotoxic ceramides and dihydroceramides [79].
FATP2 has also been implicated in the pathogenesis of non-diabetic kidney diseases. In a mouse ureteral obstruction model, FATP2-dependent FA uptake was enhanced, leading to cytokine release and fibrogenesis [80]. In mouse models of zoledronate-induced renal toxicity, and in vitro cell culture confirmation, increased proximal tubule saturated FA accumulation and fibrosis were mediated by a TGFβ-dependent increase in FATP2 expression [81]. In a study of renal transplant biopsies, enhanced FATP2 mRNA expression was associated with calcineurin inhibitor-induced interstitial fibrosis and tubular atrophy [82].
Among FA transporters not in the FATP family, Kidney Injury Molecule-1 (KIM-1), also known as T cell immunoglobulin mucin domain (TIM)-1, has recently been implicated in the pathogenesis of proximal tubule lipotoxicity. KIM-1 is expressed within the apical proximal tubule membrane and serves as a scavenger receptor for oxidized lipoproteins, as well as for apoptotic cells, through recognition of externalized phosphatidylserine [83] (Figure 2). Apical proximal tubule KIM-1 expression is increased with tubular injury [84], and constitutive overexpression caused interstitial fibrosis and GFR decline [85]. Clinically, plasma KIM-1 is a reliable biomarker for DKD [86,87,88]. In vitro and DKD mouse model studies revealed that KIM-1 mediated uptake of albumin-bound palmitate by proximal tubules [89]. Whether KIM-1 directly or indirectly transports FA was not established [89]. KIM-1-regulated FA uptake in a DKD model caused significant albuminuria, tubular atrophy, interstitial fibrosis, glomerulosclerosis, as well as secretion of inflammatory and fibrosis-inducing cytokines [89]. KIM-1-related effects on GFR were not evaluated. Consistent results were observed in the obese Zucker rat model of DKD, where cDNA microarray revealed that KIM-1 was the most highly upregulated transcript [90].
Endocannabinoids are small lipids derived from plasma membrane phospholipids, which bind in a paracrine fashion to G protein-coupled CB1 and CB2 receptors, and mediate opposing effects. Although both receptor types, as well as ligand-degrading enzymes are expressed throughout the nephron, CB1 predominates in the proximal tubule and podocytes [91,92]. In mouse and rat models of DKD, CB1 expression is upregulated in podocytes, and associated with albuminuria [93,94]. In the proximal tubule, palmitate induced proximal tubule CB1 expression, and potentiated apoptosis, suggesting that CB1 could mediate proximal tubule lipotoxicity [95]. In mice on a high fat diet, proximal tubule-specific deletion of CB1 ameliorated many aspects of DKD, including albuminuria, modest changes in GFR and interstitial fibrosis [96]. One purported mechanism was through enhanced AMP-activated protein kinase (AMPK)-mediated autophagy and enhanced FA β-oxidation [96]. These studies have led to multiple pre-clinical trials with CB1 inhibitors for DKD (see Table 1 and associated text). Studies with CB2 are more limited, though global CB2 deletion worsened albuminuria, serum creatinine and glomerulosclerosis in STZ-treated mice [97]. Although CB2 expression has been noted in cultured cells and rodent kidney [98], the lack of CB2 detection in human kidney [91] raises the issue whether the global CB2 knockout renal phenotype could be derived from extrarenal effects.
In microarray experiments from human diabetic kidneys, proximal tubule CD36 was highly expressed, stimulated by glucose, and regulated palmitate-induced apoptosis [99]. However, kidney CD36 mRNA expression was decreased or absent in mouse models of DKD [89,100,101]. To rectify these discrepancies kidney-specific CD36 transgenic mice were induced to develop diabetes with streptozotocin [33]. Although renal function was not assessed, interstitial fibrosis was not observed in the diabetic, CD36 transgenic mice. The absence of pathology was despite proximal tubule cytosolic LD accumulation, suggesting that CD36 may regulate FA compartmentalization to LDs, thereby shielding cells from lipotoxicity. Taken together, the data indicate that CD36 does not mediate proximal tubule lipotoxicity or tubular atrophy in DKD.

2.4. Role of Lipid Droplets

Numerous pathways buffer against FA accumulation and subsequent lipotoxicity. The initial response to FA uptake is metabolism by β-oxidation to generate ATP [66]. An additional defense against lipotoxicity is the Them/ACOT family of thioesterases, which accelerate catabolism of LC-CoAs, the initial FA metabolite, by hydrolyzing the thioester bond to form CoA and the corresponding FA [102,103]. Once energy needs are met, excess FAs and cholesterol are compartmentalized as TAG and cholesterol ester aggregates within the LD core [104], which is surrounded by a single phospholipid membrane. The LD membrane is decorated with proteins that facilitate LD formation, such as diacylglycerol transferase-2 (DGAT2) [105], and maintenance, such as perilipins [106], which block TAG release by inhibiting adipose triglyceride lipase (ATGL).
While all cells are capable of incorporating TAG into LDs [107], which primarily function as reservoirs for energy demands by facilitating bidirectional FA shuttling, LDs are not prominent in the cytosol of epithelial cells under normal circumstances. It is therefore reasonable to equate easily detectable LDs with a pathologic state. However, it is important to note that LD expansion represents a compensatory, protective pathway in most organs, including kidney and kidney model systems [108,109,110].
The simplest view is that LDs serve as depots for excess lipids, and thereby shield cells from “free” cholesterol and FA metabolites, by preventing lipotoxicity and lipoapoptosis [108,111,112] or by enhancing autophagy pathways [113]. A more nuanced view is predicated upon LD proximity to other organelles, which are dynamically motile within the cytosol [114]. LDs tethered to elongated mitochondria that have undergone fusion, results in enhanced generation of ATP, which fuels lipogenesis and LD expansion [115,116]. This process involves metabolic reprogramming to utilize glucose/pyruvate as a substrate, although a recent report suggests that the mitochondrial-LD juxtaposition merely enhances FA β-oxidation efficiency [117].
Two recent reports from the Pollak group revealed that FA-induced LD expansion in podocytes led to preferential translocation of APOL1 G1 and G2 risk variants from ER to LDs, which was associated with reduced cytotoxicity [118,119]. Although these studies illuminate an intriguing cytoprotective role of LDs in podocytes, it should be noted that APOL1 risk variants have been associated primarily with non-DKD [120].
Because kidney LDs have been associated with DKD (and other proteinuric renal diseases), a common strategy has been to reduce LD mass through enhanced FA β-oxidation. However, this approach risks forced mitochondrial entry of FA, which could lead to oxidative stress. Alternatively, enhanced FA sequestration into LDs could be beneficial, but unlike adipocytes, which have tremendous capacity to accommodate LDs in the cytosol, epithelial cells have limits.

3. Potential Lipotoxicity Treatments

3.1. Drugs in Current Clinical Use

3.1.1. Fibrates

Table 1. Candidate drugs for the treatment of renal lipotoxicity.
Table 1. Candidate drugs for the treatment of renal lipotoxicity.
DrugTargetMechanismReferences
FenofibratePPARαFA uptake and metabolism[121,122,123,124,125,126]
StatinsHMG-CoA reductaseCholesterol synthesis[127,128,129,130,131]
SGLT2 inhibitorsSGLT2Unknown[132,133,134,135,136,137,138,139,140]
LiraglutideGLP1 receptorsUnknown[141]
A30ABCA1Cholesterol efflux[43]
LipofermataFATP2FA uptake[80]
PBI-4050GPR40/GPR84FA uptake[142,143]
TW-37KIM-1FA uptake[89]
AM-251, rimonabantCB1FA metabolism[93,144,145,146]
INT-767FXR/TGR5FA synthesis[147]
AdipoRonAdipoR1/R2FA metabolism[148]
DehydrozingeroneMultipleFA synthesis[149]
ResveratrolSIRT1, AMPKFA synthesis and metabolism[150,151]
Several studies, in both animal models and humans, have examined the effect of fenofibrate on DKD progression. The rationale is that fibrates are activators of PPARα, which regulates FA catabolism at several levels, and theoretically reduces the propensity for lipotoxicity. In vitro studies support that decreased PPARα expression is associated with lipoapoptosis and fibrogenesis [152]. In experiments with C57BLKS Leprdb/db mice fenofibrate improved glucose tolerance and oxidative stress, as well as albuminuria and glomerular pathology [153,154]. In a non-DKD (folic acid nephropathy) model, fenofibrate normalized serum creatinine [33]. The fenofibrate effects may be partly mediated through AMPK activation and enhanced autophagy [155].
However, interpretation of fenofibrate human clinical trials is more complicated, and results are not as positive as with animal studies. In two large diabetes trials, FIELD and ACCORD, randomization to fenofibrate improved albuminuria, but caused an initial small increase in creatinine, which persisted throughout the study [121,122], and was reversible upon fenofibrate discontinuation [123]. A follow-up trial in a FIELD subcohort showed a fenofibrate-associated increase in both serum creatinine and cystatin C (indicating that the creatinine changes are not artifactual, and represent true changes in GFR), as well as no benefit in albuminuria reduction [124]. The mechanism for the initial creatinine increase is unclear, but may be related to fenofibrate-induced inhibition of vasodilator prostaglandins [125]. Despite the initial effects on GFR, the rate of eGFR decline over the course of the ACCORD trial was significantly lower in the group randomized to fenofibrate [126]. Newer PPARα activators, which are not metabolized by kidney, may represent safer alternatives to treat renal lipotoxicity [156].
One explanation for the discrepancy of the animal and human results is that PPARα affects the transcription of multiple lipid metabolism genes, both protective and deleterious (e.g., FATP2). In addition, effects on GFR are strain-dependent in mice, and may not be manifest in mouse DKD models with forgiving genetic backgrounds.

3.1.2. HMG-CoA Reductase Inhibitors (Statins)

Statins are among the most widely prescribed drugs in the world, with demonstrated cardiovascular benefits in patients with CKD [127], though the effects are attenuated with lower baseline eGFRs [128]. The effect of statins specifically on DKD is less clear. Two recent meta-analyses of randomized controlled trials showed that statins were associated with modest reduction of albuminuria, but not with modulation of eGFR decline [129,130]. These data are consistent with a larger meta-analysis of DKD and non-DKD patients, which noted no effect on eGFR in the DKD subgroup analysis [131].

3.1.3. Proprotein Convertase Subtilisin-Kexin Type 9 (PCSK9) Inhibitors

Inhibitors of PCSK9 (evolocumab and alirocumab) are now recognized as alternatives or adjuncts to statins for LDL-cholesterol reduction. The PCSK9 gene is expressed at very low levels in the kidney, with modest modulation in the loop of Henle and collecting duct with diabetes [101]. A recent review concludes that PCSK9 inhibitors are safe and effective for cholesterol reduction with modest CKD (including DKD), though similar data are unavailable for patients with eGFR < 30 mL/min/1.73 m2 [157]. There are also insufficient data regarding the role of PCSK9 inhibitors in renal lipotoxicity or DKD progression.

3.1.4. Sodium-Glucose Transporter-2 Inhibitors (SGLT2i)

SGLT2i are now recognized as the standard of care for patients with DKD and eGFR >25 mL/min/1.73 m2, and adjunctive to glycemic control and renin-angiotensin-aldosterone antagonist therapies [132,133,134]. However, regiments that include SGLT2i-induced prevention of DKD progression is incompletely established, and likely to be multi-factorial, including some evidence for anti-lipotoxic effects [135]. Dapagliflozin prevented renal tubule lipid accumulation and fibrosis in mice maintained on a high fat diet [136]. In the FLS-ob/ob mouse model of diabetes, ipragliflozin also reduced tubular lipid deposition, and was associated with decreased apoptosis and ER stress [137]. Comparable results were observed with dapagliflozin in Leprdb/db mice [138]. Two additional studies in high fat diet-fed mice demonstrated that SGLT2i preserved renal tubule mitochondrial morphology and function [139,140].

3.1.5. Glucagon-like Peptide-1 (GLP-1) Receptor Agonists

Like SGLT2 inhibitors liraglutide has been shown to have renoprotective effects in clinical trials, but the benefit is mainly limited to reduction in proteinuria. In preclinical studies, a 12-week course of liraglutide improved plasma lipid profiles and reduced proximal tubule lipid droplets in an inducible rat model of DKD [141], suggesting that it might prevent lipotoxicity and renal dysfunction.

3.1.6. Mineralocorticoid Receptor Antagonists (MRA)

Inhibition of the renin-angiotensin-aldosterone cascade has been a standard treatment for DKD for decades. While the pathophysiologic focus has largely been on angiotensin II, aldosterone has independent effects, which include inflammation and fibrosis [158]. Recent studies have shown that the non-steroidal MRA, finerenone, may be benefical when added to renin-angiotensin inhibitors, for slowing the progression of DKD [159,160,161]. There are multiple reports of spironolactone and eplerenone efficacy for hepatosteatosis and liver fibrosis in mouse models [162,163,164], but data are lacking regarding the impact of MRA on renal lipotoxicity.

3.2. Experimental Compounds

3.2.1. ABCA1 Induction

In conjunction with experiments to characterize the role of ABCA1 in podocyte protection, Ducasa et al. employed a previously described inducer of ABCA1 expression (A30) in their in vivo studies [43]. Leprdb/db mice treated with A30 for two to four weeks demonstrated restoration of ABCA1 expression, as well as reductions in albuminuria and BUN. Histologic analyses revealed decreased cortical cholesterol, FA and peroxidized cardiolipin accumulation, and improved glomerular histopathology [43]. Other small molecule (pyridine carboxamide) inducers of ABCA1 are under investigation as potential therapies for DKD, as well as for FSGS and Alport’s syndrome.
Sterol-O-acyltransferase-1 (SOAT1) catalyzes the esterification of cholesterol to cholesterol esters, which are then stored in LDs. Pharmacologic SOAT1 inhibition with Sandoz 58-035 improved albuminuria, serum creatinine and podocyte LD accumulation in a mouse model of Alport’s syndrome [165]. In vitro studies in culture podocytes revealed that SOAT1 inhibition was cytoprotective, by a mechanism that involved enhanced ABCA1 expression [165].

3.2.2. FATP2 Inhibitors

Two small molecular weight FATP2 inhibitors (lipofermata and grassofermata) were identified from high throughput screens, with low μM IC50 in fluorescent FA uptake assays [166]. Although pharmacokinetic data are limited, in a mouse model of urinary tract obstruction, oral administration of lipofermata prevented tubule lipid accumulation and fibrosis; renal function studies were not conducted [80].

3.2.3. G-protein-Coupled FA Receptor Inhibitors

GP40 (also known as FFAR1) and GP84 are G-protein-couple FA receptors expressed predominantly in podocytes in kidney [14], though mRNA is detectable in proximal tubule and collecting duct [142]. PBI-4050 is a synthetic medium-chain fatty acid analog with agonist and antagonist activity toward GPR40 and GPR84, respectively. In experiments with the Leprdb/db eNOS-/- model of type 2 diabetes and progressive kidney disease, PBI-4050 was effective in primary prevention and as treatment for established DKD, as determined by reductions in albuminuria, GFR decline and glomerulosclerosis [142]. Interestingly, PBI-4050 also improved glucose tolerance by preserving pancreatic β-cell function and insulin secretion [142], similar to the phenotype observed in FATP2 knockout mice [14]. Additional studies in GPR40 and GPR84 knockout mice crossed with non-diabetic models of interstitial fibrosis revealed that GPR40 was protective, whereas GPR84 was deleterious. PBI-4050 also inhibited fibrosis in these non-diabetic models [143].
GPR43 (also known as FFAR2) is a short-chain FA receptor. In mice with type 1 diabetes induced by streptozotocin, superimposed GPR43 deletion attenuated LDL receptor-mediated cholesterol accumulation in podocytes [167]. Further studies in cultured podocyte identified autophagy inhibition and ERK-dependent activation of the EGR1 transcription factor as potential mechanisms. Treatment of cultured podocytes with the ERK 1/2 inhibitor PD0325901 decreased LDL receptor expression and cholesterol accumulation, and enhanced autophagy [167].

3.2.4. KIM-1 Inhibition

Mori et al. conducted a high throughput screen of a 14,430 compound small molecule library, using inhibition of oxidized LDL uptake as the readout in high KIM-1 expressing kidney-derived 769-P cells [89]. These experiments yielded TW-37 as a candidate KIM-1 inhibitor. Extensive in vitro and in vivo (non-DKD, palmitate complexed to albumin overload model) studies revealed that TW-37 inhibited KIM-1-mediated FA uptake, and blunted interstitial inflammation and fibrosis [89].

3.2.5. CB1 Inhibition

AM-251 and rimonabant are structurally similar biarylpyrazole compounds with inverse agonist/antagonist activity for the CB1 cannabinoid receptor. In animal models, rimonabant reduced albuminuria and glomerulosclerosis in obese fa/fa Zucker rats [144], and decreased albuminuria and kidney lipid content in Leprdb/db mice [145]. AM-251 also reduced albuminuria in streptozotocin-treated mice and a rat model of obesity [93,146]. Jourdan et al. demonstrated improved albuminuria and GFR in ZDF rats with another CB1 inverse agonist, JD5037 [94]. Rimonabant is being tested for effects on glycemia and obesity in several clinical trials, though none are measuring DKD outcomes.

3.2.6. FXR/TGR5 Activation

The FXR nuclear hormone and TGR5 G protein–coupled bile acid receptors have been implicated in the pathogenesis of DKD [55,168]. RNAseq analysis (and targeted protein expression confirmation) of kidneys from streptozotocin-induced diabetic mice revealed that treatment with the selective FXR (INT-747) or TGR5 (INT-777) agonists affect distinct pathways [147]; INT-747 uniquely increased SREBP-1-mediated lipogenesis, whereas INT-777 enhanced mitochondrial biogenesis. Importantly, the combined FXR/TGR5 agonist (INT-767) decreased cortical FA and cholesterol accumulation, and dramatically reduced albuminuria, as well as interstitial and glomerular fibrosis in both the streptozotocin and Leprdb/db diabetic mouse models [147].

3.2.7. Adiponectin Receptor Activation

Adiponectin is a cytokine that is secreted by adipocytes, and has been associated with amelioration of diabetic effects through anti-inflammatory, anti-fibrotic, and anti-oxidant activities. Adiponectin regulates FA metabolism by inducing AMPK phosphorylation and increasing peroxisome PPARα expression. AdipoR1 and AdipoR2 adiponectin receptors are expressed throughout the human kidney, but glomerular expression was decreased in DKD, coinciding with early GFR decline [148]. AdipoRon is a synthetic adiponectin receptor agonist which binds to both AdipoR1 and AdipoR2 receptors [169]. AdipoRon treatment for four weeks significantly reduced albuminuria in Leprdb/db mice, and was associated with enhanced Ca2+/calmodulin-dependent protein kinase, liver kinase B1 and AMPK enzyme activities, as well as increased PPARα expression [148].

3.2.8. Dehydrozingerone

Dehydrozingerone (DHZ) is a curcumin analog that exhibits anti-neoplastic, anti-oxidant and anti-diabetic effects. C57BL/6 mice on a high fat diet and treated with DHZ for four weeks experienced significant reductions in albuminuria, as well as reduced kidney FA and cholesterol content, and improved glomerular histology [149]. Expression patterns derived from quantitative PCR analysis of kidney samples suggested several mechanisms for DHZ efficacy, including reduced FA synthesis and decreased oxidative stress.

3.2.9. Resveratrol

Resveratrol is a polyphenol compound that is abundant in grapes and red wine, which can also be taken as a dietary supplement. Purported metabolic benefits are myriad, including reduced serum glucose and lipid concentrations. Pre-clinical studies suggest resveratrol regulates multiple mechanisms, including reduction in FA synthesis, inflammation, oxidative and ER stress, mitochondrial dysfunction, autophagy and apoptosis [170]. In a recent study with high fat diet-fed C57BL/6J mice, resveratrol supplementation decreased blood glucose and lipid concentrations, and was associated with a trend toward lower serum creatinine [150]. However, a recent meta-analysis of randomized clinical diabetes trials showed a significant benefit of resveratrol for lower systolic blood pressure, but no effect on serum creatinine, though there was a paucity of studies evaluating relevant DKD biomarkers [151].

4. Conclusions

Lipotoxicity is an increasingly recognized mechanism of DKD pathophysiology in the podocyte and proximal tubule. Because multiple pathways have been identified, and the mechanisms of podocyte and proximal tubule lipotoxicity may differ (Figure 1 and Figure 2), there may be opportunities to cultivate multiple kidney therapeutic targets, e.g., different glomerular and tubular therapies [171]. This strategy could conceivably result in therapeutic synergism, much like the combined effects of ACE inhibitor or ARBs with SGLT2 inhibitors for DKD, or chemotherapy protocols.

Funding

This research was funded by NIH grant 2R01 DK067528.

Acknowledgments

Thanks to Jeffrey Garvin for critical review of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, L.O.; Klett, E.L.; Coleman, R.A. Acyl-CoA synthesis, lipid metabolism and lipotoxicity. Biochim. Biophys. Acta 2010, 1801, 246–251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Garbarino, J.; Sturley, S.L. Saturated with fat: New perspectives on lipotoxicity. Curr. Opin. Clin. Nutr. Metab. Care 2009, 12, 110–116. [Google Scholar] [CrossRef] [PubMed]
  3. Schaffer, J.E. Lipotoxicity: When tissues overeat. Curr. Opin. Lipidol. 2003, 14, 281–287. [Google Scholar] [CrossRef] [PubMed]
  4. Musso, G.; Gambino, R.; Tabibian, J.H.; Ekstedt, M.; Kechagias, S.; Hamaguchi, M.; Hultcrantz, R.; Hagström, H.; Yoon, S.K.; Charatcharoenwitthaya, P.; et al. Association of non-alcoholic fatty liver disease with chronic kidney disease: A systematic review and meta-analysis. PLoS Med. 2014, 11, e1001680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Sheka, A.C.; Adeyi, O.; Thompson, J.; Hameed, B.; Crawford, P.A.; Ikramuddin, S. Nonalcoholic steatohepatitis: A Review. JAMA 2020, 323, 1175–1183. [Google Scholar] [CrossRef] [PubMed]
  6. Tandra, S.; Yeh, M.M.; Brunt, E.M.; Vuppalanchi, R.; Cummings, O.W.; Ünalp-Arida, A.; Wilson, L.A.; Chalasani, N. Presence and significance of microvesicular steatosis in nonalcoholic fatty liver disease. J. Hepatol. 2011, 55, 654–659. [Google Scholar] [CrossRef] [Green Version]
  7. Kimmelstiel, P.; Wilson, C. Intercapillary lesions in the glomeruli of the kidney. Am. J. Pathol. 1936, 12, 83–98. [Google Scholar]
  8. Oliver, J.; MacDowell, M.; Lee, Y.C. Cellular mechanisms of protein metabolism in the nephron. I. The structural aspects of proteinuria; tubular absorption, droplet formation, and the disposal of proteins. J. Exp. Med. 1954, 99, 589–604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Moorhead, J.F.; Chan, M.K.; El-Nahas, M.; Varghese, Z. Lipid nephrotoxicity in chronic progressive glomerular and tubulo-interstitial disease. Lancet 1982, 2, 1309–1311. [Google Scholar] [CrossRef]
  10. Weinberg, J.M. Lipotoxicity. Kidney Int. 2006, 70, 1560–1566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Ruan, X.Z.; Varghese, Z.; Moorhead, J.F. An update on the lipid nephrotoxicity hypothesis. Nat. Rev. Nephrol. 2009, 5, 713–721. [Google Scholar] [CrossRef] [PubMed]
  12. Abu Jawdeh, B.G.; Khan, S.; Deschenes, I.; Hoshi, M.; Goel, M.; Lock, J.T.; Shinlapawittayatorn, K.; Babcock, G.; Lakhe-Reddy, S.; DeCaro, G.; et al. Phosphoinositide binding differentially regulates NHE1 Na+/H+ exchanger-dependent proximal tubule cell survival. J. Biol. Chem. 2011, 286, 42435–42445. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Khan, S.; Abu Jawdeh, B.G.; Goel, M.; Schilling, W.P.; Parker, M.D.; Puchowicz, M.A.; Yadav, S.P.; Harris, R.C.; El-Meanawy, A.; Hoshi, M.; et al. Lipotoxic disruption of NHE1 interaction with PI(4,5)P2 expedites proximal tubule apoptosis. J. Clin. Invest. 2014, 124, 1057–1068. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Khan, S.; Gaivin, R.J.; Abramovich, C.; Boylan, M.; Calles, J.; Schelling, J.R. Fatty acid transport protein-2 (FATP2) regulates glycemic control and diabetic kidney disease progression. JCI insight 2020, 5, e136845. [Google Scholar] [CrossRef] [PubMed]
  15. Riedel, M.J.; Light, P.E. Saturated and cis/trans unsaturated acyl CoA esters differentially regulate wild-type and polymorphic beta-cell ATP-sensitive K+ channels. Diabetes 2005, 54, 2070–2079. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Riedel, M.J.; Baczko, I.; Searle, G.J.; Webster, N.; Fercho, M.; Jones, L.; Lang, J.; Lytton, J.; Dyck, J.R.; Light, P.E. Metabolic regulation of sodium-calcium exchange by intracellular acyl CoAs. EMBO J. 2006, 25, 4605–4614. [Google Scholar] [CrossRef] [PubMed]
  17. Reaven, G.M.; Hollenbeck, C.; Jeng, C.Y.; Wu, M.S.; Chen, Y.D. Measurement of plasma glucose, free fatty acid, lactate, and insulin for 24 h in patients with NIDDM. Diabetes 1988, 37, 1020–1024. [Google Scholar] [CrossRef] [PubMed]
  18. Afshinnia, F.; Rajendiran, T.M.; Soni, T.; Byun, J.; Wernisch, S.; Sas, K.M.; Hawkins, J.; Bellovich, K.; Gipson, D.; Michailidis, G.; et al. Impaired β-oxidation and altered complex lipid fatty acid partitioning with advancing CKD. J. Am. Soc. Nephrol. 2018, 29, 295–306. [Google Scholar] [CrossRef] [Green Version]
  19. Listenberger, L.L.; Han, X.; Lewis, S.E.; Cases, S.; Farese, R.V., Jr.; Ory, D.S.; Schaffer, J.E. Triglyceride accumulation protects against fatty acid-induced lipotoxicity. Proc. Natl. Acad. Sci. U.S.A. 2003, 100, 3077–3082. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Pettus, B.J.; Chalfant, C.E.; Hannun, Y.A. Ceramide in apoptosis: An overview and current perspectives. Biochim. Biophys. Acta 2002, 1585, 114–125. [Google Scholar] [CrossRef]
  21. Aflaki, E.; Radovic, B.; Chandak, P.G.; Kolb, D.; Eisenberg, T.; Ring, J.; Fertschai, I.; Uellen, A.; Wolinski, H.; Kohlwein, S.D.; et al. Triacylglycerol accumulation activates the mitochondrial apoptosis pathway in macrophages. J. Biol. Chem. 2011, 286, 7418–7428. [Google Scholar] [CrossRef] [PubMed]
  22. Epstein, S.; Kirkpatrick, C.L.; Castillon, G.A.; Muniz, M.; Riezman, I.; David, F.P.; Wollheim, C.B.; Riezman, H. Activation of the unfolded protein response pathway causes ceramide accumulation in yeast and INS-1E insulinoma cells. J. Lipid Res. 2012, 53, 412–420. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Han, J.; Kaufman, R.J. The role of ER stress in lipid metabolism and lipotoxicity. J. Lipid Res. 2016, 57, 1329–1338. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. van Timmeren, M.M.; Gross, M.L.; Hanke, W.; Klok, P.A.; van, G.H.; Stegeman, C.A.; Bakker, S.J. Oleic acid loading does not add to the nephrotoxic effect of albumin in an amphibian and chronic rat model of kidney injury. Nephrol.Dial.Transplant. 2008, 23, 3814–3823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Sieber, J.; Lindenmeyer, M.T.; Kampe, K.; Campbell, K.N.; Cohen, C.D.; Hopfer, H.; Mundel, P.; Jehle, A.W. Regulation of podocyte survival and endoplasmic reticulum stress by fatty acids. Am. J. Physiol. Renal Physiol. 2010, 229, F821–F829. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Thomas, M.E.; Morrison, A.R.; Schreiner, G.F. Metabolic effects of fatty acid-bearing albumin on a proximal tubule cell line. Am. J. Physiol. 1995, 268, F1177–F1184. [Google Scholar] [CrossRef] [PubMed]
  27. Iwai, T.; Kume, S.; Chin-Kanasaki, M.; Kuwagata, S.; Araki, H.; Takeda, N.; Sugaya, T.; Uzu, T.; Maegawa, H.; Araki, S.I. Stearoyl-CoA desaturase-1 protects cells against lipotoxicity-mediated apoptosis in proximal tubular cells. Int. J. Mol. Sci. 2016, 17, 1868. [Google Scholar] [CrossRef] [Green Version]
  28. Lee, H.S.; Suh, J.Y.; Kang, B.C.; Lee, Y.J. Lipotoxicity dysregulates immunoproteasome in podocytes and kidneys of type 2 diabetes. Am. J. Physiol. Renal Physiol. 2021, 320, F548–F558. [Google Scholar] [CrossRef]
  29. Proctor, G.; Jiang, T.; Iwahashi, M.; Wang, Z.; Li, J.; Levi, M. Regulation of renal fatty acid and cholesterol metabolism, inflammation, and fibrosis in Akita and OVE26 mice with type 1 diabetes. Diabetes 2006, 55, 2502–2509. [Google Scholar] [CrossRef] [Green Version]
  30. Chung, J.J.; Huber, T.B.; Godel, M.; Jarad, G.; Hartleben, B.; Kwoh, C.; Keil, A.; Karpitskiy, A.; Hu, J.; Huh, C.J.; et al. Albumin-associated free fatty acids induce macropinocytosis in podocytes. J. Clin. Invest. 2015, 125, 2307–2316. [Google Scholar] [CrossRef] [Green Version]
  31. Khan, S.; Cabral, P.D.; Schilling, W.P.; Schmidt, Z.W.; Uddin, A.N.; Gingras, A.; Madhavan, S.M.; Garvin, J.L.; Schelling, J.R. Kidney proximal tubule lipoapoptosis Is regulated by Fatty Acid Transporter-2 (FATP2). J. Am. Soc. Nephrol. 2018, 29, 81–91. [Google Scholar] [CrossRef] [PubMed]
  32. Merscher-Gomez, S.; Guzman, J.; Pedigo, C.E.; Lehto, M.; Aguillon-Prada, R.; Mendez, A.; Lassenius, M.I.; Forsblom, C.; Yoo, T.; Villarreal, R.; et al. Cyclodextrin protects podocytes in diabetic kidney disease. Diabetes 2013, 62, 3817–3827. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Kang, H.M.; Ahn, S.H.; Choi, P.; Ko, Y.A.; Han, S.H.; Chinga, F.; Park, A.S.; Tao, J.; Sharma, K.; Pullman, J.; et al. Defective fatty acid oxidation in renal tubular epithelial cells has a key role in kidney fibrosis development. Nat. Med. 2015, 21, 37–46. [Google Scholar] [CrossRef] [PubMed]
  34. Rosca, M.G.; Vazquez, E.J.; Chen, Q.; Kerner, J.; Kern, T.S.; Hoppel, C.L. Oxidation of fatty acids is the source of increased mitochondrial reactive oxygen species production in kidney cortical tubules in early diabetes. Diabetes 2012, 61, 2074–2083. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Yoon, H.; Shaw, J.L.; Haigis, M.C.; Greka, A. Lipid metabolism in sickness and in health: Emerging regulators of lipotoxicity. Mol. Cell 2021, 81, 3708–3730. [Google Scholar] [CrossRef]
  36. Lipke, K.; Kubis-Kubiak, A.; Piwowar, A. Molecular mechanism of lipotoxicity as an interesting aspect in the development of pathological states-Current view of knowledge. Cells 2022, 11, 844. [Google Scholar] [CrossRef]
  37. Schwarz, K.; Simons, M.S.; Reiser, J.; Saleem, M.A.; Faul, C.; Kriz, W.; Shaw, A.S.; Holzman, L.B.; Mundel, P. Podocin, a raft-associated component of the glomerular slit diaphragm, interacts with CD2AP and nephrin. J. Clin. Invest. 2001, 108, 1621–1629. [Google Scholar] [CrossRef]
  38. Huber, T.B.; Simons, M.; Hartleben, B.; Sernetz, L.; Schmidts, M.; Gundlach, E.; Saleem, M.A.; Walz, G.; Benzing, T. Molecular basis of the functional podocin-nephrin complex: Mutations in the NPHS2 gene disrupt nephrin targeting to lipid raft microdomains. Hum. Mol. Genet. 2003, 12, 3397–3405. [Google Scholar] [CrossRef] [Green Version]
  39. Yoo, T.H.; Pedigo, C.E.; Guzman, J.; Correa-Medina, M.; Wei, C.; Villarreal, R.; Mitrofanova, A.; Leclercq, F.; Faul, C.; Li, J.; et al. Sphingomyelinase-like phosphodiesterase 3b expression levels determine podocyte injury phenotypes in glomerular disease. J. Am. Soc. Nephrol. 2014, 26, 133–147. [Google Scholar] [CrossRef] [Green Version]
  40. Mitrofanova, A.; Mallela, S.K.; Ducasa, G.M.; Yoo, T.H.; Rosenfeld-Gur, E.; Zelnik, I.D.; Molina, J.; Varona Santos, J.; Ge, M.; Sloan, A.; et al. SMPDL3b modulates insulin receptor signaling in diabetic kidney disease. Nat. Commun. 2019, 10, 2692. [Google Scholar] [CrossRef] [Green Version]
  41. Herman-Edelstein, M.; Scherzer, P.; Tobar, A.; Levi, M.; Gafter, U. Altered renal lipid metabolism and renal lipid accumulation in human diabetic nephropathy. J. Lipid Res. 2014, 55, 561–572. [Google Scholar] [CrossRef]
  42. Pedigo, C.E.; Ducasa, G.M.; Leclercq, F.; Sloan, A.; Mitrofanova, A.; Hashmi, T.; Molina-David, J.; Ge, M.; Lassenius, M.I.; Forsblom, C.; et al. Local TNF causes NFATc1-dependent cholesterol-mediated podocyte injury. J. Clin. Invest. 2016, 126, 3336–3350. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Ducasa, G.M.; Mitrofanova, A.; Mallela, S.K.; Liu, X.; Molina, J.; Sloan, A.; Pedigo, C.E.; Ge, M.; Santos, J.V.; Hernandez, Y.; et al. ATP-binding cassette A1 deficiency causes cardiolipin-driven mitochondrial dysfunction in podocytes. J. Clin. Invest. 2019, 129, 3387–3400. [Google Scholar] [CrossRef] [Green Version]
  44. Brosius, F.C., 3rd; Alpers, C.E.; Bottinger, E.P.; Breyer, M.D.; Coffman, T.M.; Gurley, S.B.; Harris, R.C.; Kakoki, M.; Kretzler, M.; Leiter, E.H.; et al. Mouse models of diabetic nephropathy. J. Am. Soc. Nephrol. 2009, 20, 2503–2512. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Hudkins, K.L.; Pichaiwong, W.; Wietecha, T.; Kowalewska, J.; Banas, M.C.; Spencer, M.W.; Muhlfeld, A.; Koelling, M.; Pippin, J.W.; Shankland, S.J.; et al. BTBR Ob/Ob mutant mice model progressive diabetic nephropathy. J. Am. Soc. Nephrol. 2010, 21, 1533–1542. [Google Scholar] [CrossRef] [Green Version]
  46. Chen, N.; Mu, L.; Yang, Z.; Du, C.; Wu, M.; Song, S.; Yuan, C.; Shi, Y. Carbohydrate response element-binding protein regulates lipid metabolism via mTOR complex1 in diabetic nephropathy. J. Cell. Physiol. 2020, 236, 625–640. [Google Scholar] [CrossRef] [PubMed]
  47. Fu, Y.; Sun, Y.; Wang, M.; Hou, Y.; Huang, W.; Zhou, D.; Wang, Z.; Yang, S.; Tang, W.; Zhen, J.; et al. Elevation of JAML promotes diabetic kidney disease by modulating podocyte lipid metabolism. Cell Metab. 2020, 32, 1052–1062. [Google Scholar] [CrossRef] [PubMed]
  48. Sieber, J.; Weins, A.; Kampe, K.; Gruber, S.; Lindenmeyer, M.T.; Cohen, C.D.; Orellana, J.M.; Mundel, P.; Jehle, A.W. Susceptibility of Podocytes to Palmitic Acid Is Regulated by Stearoyl-CoA Desaturases 1 and 2. Am. J. Pathol. 2013, 783, 735–744. [Google Scholar] [CrossRef] [Green Version]
  49. Clement, L.C.; Mace, C.; Avila-Casado, C.; Joles, J.A.; Kersten, S.; Chugh, S.S. Circulating angiopoietin-like 4 links proteinuria with hypertriglyceridemia in nephrotic syndrome. Nat. Med. 2014, 20, 37–46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Hua, W.; Huang, H.Z.; Tan, L.T.; Wan, J.M.; Gui, H.B.; Zhao, L.; Ruan, X.Z.; Chen, X.M.; Du, X.G. CD36 mediated fatty ccid-induced podocyte apoptosis via oxidative stress. PLoS ONE 2015, 10, e0127507. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Falkevall, A.; Mehlem, A.; Palombo, I.; Heller Sahlgren, B.; Ebarasi, L.; He, L.; Ytterberg, A.J.; Olauson, H.; Axelsson, J.; Sundelin, B.; et al. Reducing VEGF-B signaling ameliorates renal lipotoxicity and protects against diabetic kidney disease. Cell Metab. 2017, 25, 713–726. [Google Scholar] [CrossRef] [PubMed]
  52. Sun, L.; Halaihel, N.; Zhang, W.; Rogers, T.; Levi, M. Role of sterol regulatory element-binding protein 1 in regulation of renal lipid metabolism and glomerulosclerosis in diabetes mellitus. J. Biol. Chem. 2002, 277, 18919–18927. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Wang, Z.; Jiang, T.; Li, J.; Proctor, G.; McManaman, J.L.; Lucia, S.; Chua, S.; Levi, M. Regulation of renal lipid metabolism, lipid accumulation, and glomerulosclerosis in FVBdb/db mice with type 2 diabetes. Diabetes 2005, 54, 2328–2335. [Google Scholar] [CrossRef] [PubMed]
  54. Park, C.W.; Kim, H.W.; Ko, S.H.; Chung, H.W.; Lim, S.W.; Yang, C.W.; Chang, Y.S.; Sugawara, A.; Guan, Y.; Breyer, M.D. Accelerated diabetic nephropathy in mice lacking the peroxisome proliferator-activated receptor alpha. Diabetes 2006, 55, 885–893. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Jiang, T.; Wang, X.X.; Scherzer, P.; Wilson, P.; Tallman, J.; Takahashi, H.; Li, J.; Iwahashi, M.; Sutherland, E.; Arend, L.; et al. Farnesoid X receptor modulates renal lipid metabolism, fibrosis, and diabetic nephropathy. Diabetes 2007, 56, 2485–2493. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Wang, X.X.; Jiang, T.; Shen, Y.; Caldas, Y.; Miyazaki-Anzai, S.; Santamaria, H.; Urbanek, C.; Solis, N.; Scherzer, P.; Lewis, L.; et al. Diabetic nephropathy is accelerated by farnesoid X receptor deficiency and inhibited by farnesoid X receptor activation in a type 1 diabetes model. Diabetes 2010, 59, 2916–2927. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Dickman, K.G.; Mandel, L.J. Glycolytic and oxidative metabolism in primary renal proximal tubule cultures. Am. J. Physiol. 1989, 257, C333–C340. [Google Scholar] [CrossRef] [PubMed]
  58. Dickman, K.G.; Mandel, L.J. Differential effects of respiratory inhibitors on glycolysis in proximal tubules. Am. J. Physiol. 1990, 258, F1608–F1615. [Google Scholar] [CrossRef] [PubMed]
  59. Zager, R.A.; Johnson, A.C.; Becker, K. Renal cortical pyruvate depletion during AKI. J. Am. Soc.. Nephrol. 2014, 25, 998–1012. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Lan, R.; Geng, H.; Singha, P.K.; Saikumar, P.; Bottinger, E.P.; Weinberg, J.M.; Venkatachalam, M.A. Mitochondrial pathology and glycolytic shift during proximal tubule atrophy after ischemic AKI. J. Am. Soc. Nephrol. 2016, 27, 3356–3367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Cai, T.; Ke, Q.; Fang, Y.; Wen, P.; Chen, H.; Yuan, Q.; Luo, J.; Zhang, Y.; Sun, Q.; Lv, Y.; et al. Sodium-glucose cotransporter 2 inhibition suppresses HIF-1α-mediated metabolic switch from lipid oxidation to glycolysis in kidney tubule cells of diabetic mice. Cell Death Dis. 2020, 11, 390. [Google Scholar] [CrossRef] [PubMed]
  62. Meyer, C.; Dostou, J.; Nadkarni, V.; Gerich, J. Effects of physiological hyperinsulinemia on systemic, renal, and hepatic substrate metabolism. Am. J. Physiol. 1998, 275, F915–F921. [Google Scholar] [CrossRef] [PubMed]
  63. Sas, K.M.; Kayampilly, P.; Byun, J.; Nair, V.; Hinder, L.M.; Hur, J.; Zhang, H.; Lin, C.; Qi, N.R.; Michailidis, G.; et al. Tissue-specific metabolic reprogramming drives nutrient flux in diabetic complications. JCI insight 2016, 1, e86976. [Google Scholar] [CrossRef] [PubMed]
  64. Li, J.; Liu, H.; Takagi, S.; Nitta, K.; Kitada, M.; Srivastava, S.P.; Takagaki, Y.; Kanasaki, K.; Koya, D. Renal protective effects of empagliflozin via inhibition of EMT and aberrant glycolysis in proximal tubules. JCI insight 2020, 5, e129034. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Lewy, P.R.; Quintanilla, A.; Levin, N.W.; Kessler, R.H. Renal energy metabolism and sodium reabsorption. Annu. Rev. Med. 1973, 24, 365–384. [Google Scholar] [CrossRef]
  66. Guder, W.G.; Wagner, S.; Wirthensohn, G. Metabolic fuels along the nephron: Pathways and intracellular mechanisms of interaction. Kidney Int. 1986, 29, 41–45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Balaban, R.S.; Mandel, L.J. Metabolic substrate utilization by rabbit proximal tubule. An NADH fluorescence study. Am. J. Physiol. 1988, 254, F407–F416. [Google Scholar] [CrossRef]
  68. Kamijo, A.; Kimura, K.; Sugaya, T.; Yamanouchi, M.; Hase, H.; Kaneko, T.; Hirata, Y.; Goto, A.; Fujita, T.; Omata, M. Urinary free fatty acids bound to albumin aggravate tubulointerstitial damage. Kidney Int. 2002, 62, 1628–1637. [Google Scholar] [CrossRef] [Green Version]
  69. Arici, M.; Chana, R.; Lewington, A.; Brown, J.; Brunskill, N.J. Stimulation of proximal tubular cell apoptosis by albumin-bound fatty acids mediated by peroxisome proliferator activated receptor-γ. J. Am. Soc. Nephrol. 2003, 14, 17–27. [Google Scholar] [CrossRef] [Green Version]
  70. Hou, B.; He, P.; Ma, P.; Yang, X.; Xu, C.; Lam, S.M.; Shui, G.; Yang, X.; Zhang, L.; Qiang, G.; et al. Comprehensive lipidome profiling of the kidney in early-stage diabetic nephropathy. Frontiers Endocrinol. 2020, 11, 359. [Google Scholar] [CrossRef]
  71. Kuppe, C.; Ibrahim, M.M.; Kranz, J.; Zhang, X.; Ziegler, S.; Perales-Patón, J.; Jansen, J.; Reimer, K.C.; Smith, J.R.; Dobie, R.; et al. Decoding myofibroblast origins in human kidney fibrosis. Nature 2020, 589, 281–286. [Google Scholar] [CrossRef] [PubMed]
  72. Myers, B.D.; Winetz, J.A.; Chui, F.; Michaels, A.S. Mechanisms of proteinuria in diabetic nephropathy: A study of glomerular barrier function. Kidney Int. 1982, 21, 633–641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Ruggenenti, P.; Mosconi, L.; Sangalli, F.; Casiraghi, F.; Gambara, V.; Remuzzi, G.; Remuzzi, A. Glomerular size-selective dysfunction in NIDDM is not ameliorated by ACE inhibition or by calcium channel blockade. Kidney Int. 1999, 55, 984–994. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Robles, N.R.; Villa, J.; Gallego, R.H. Non-proteinuric diabetic nephropathy. J. Clin. Med. 2015, 4, 1761–1773. [Google Scholar] [CrossRef] [PubMed]
  75. Berhane, A.M.; Weil, E.J.; Knowler, W.C.; Nelson, R.G.; Hanson, R.L. Albuminuria and estimated glomerular filtration rate as predictors of diabetic end-stage renal disease and death. Clin. J. Am. Soc. Nephrol. 2011, 6, 2444–2451. [Google Scholar] [CrossRef] [Green Version]
  76. Ruggiero, C.; Elks, C.M.; Kruger, C.; Cleland, E.; Addison, K.; Noland, R.C.; Stadler, K. Albumin-bound fatty acids but not albumin itself alter redox balance in tubular epithelial cells and induce a peroxide-mediated redox-sensitive apoptosis. Am. J. Physiol. Renal Physiol. 2014, 306, F296–F306. [Google Scholar] [CrossRef] [PubMed]
  77. Lowell, B.B.; Shulman, G.I. Mitochondrial dysfunction and type 2 diabetes. Science 2005, 307, 384–387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Reidy, K.; Kang, H.M.; Hostetter, T.; Susztak, K. Molecular mechanisms of diabetic kidney disease. J. Clin. Invest. 2014, 124, 2333–2340. [Google Scholar] [CrossRef]
  79. Kim, J.L.; Mestre, B.; Malitsky, S.; Itkin, M.; Kupervaser, M.; Futerman, A.H. Fatty acid transport protein 2 interacts with ceramide synthase 2 to promote ceramide synthesis. J. Biol. Chem. 2022, 298, 101735. [Google Scholar] [CrossRef]
  80. Chen, Y.; Yan, Q.; Lv, M.; Song, K.; Dai, Y.; Huang, Y.; Zhang, L.; Zhang, C.; Gao, H. Involvement of FATP2-mediated tubular lipid metabolic reprogramming in renal fibrogenesis. Cell Death Dis. 2020, 11, 994. [Google Scholar] [CrossRef]
  81. Cheng, L.; Ge, M.; Lan, Z.; Ma, Z.; Chi, W.; Kuang, W.; Sun, K.; Zhao, X.; Liu, Y.; Feng, Y.; et al. Zoledronate dysregulates fatty acid metabolism in renal tubular epithelial cells to induce nephrotoxicity. Arch. Toxicol. 2018, 92, 469–485. [Google Scholar] [CrossRef]
  82. Maluf, D.G.; Dumur, C.I.; Suh, J.L.; Lee, J.K.; Cathro, E.P.; King, A.L.; Gallon, L.; Brayman, K.L.; Mas, V.R. Evaluation of molecular profiles in calcineurin inhibitor toxicity post-kidney transplant: Input to chronic allograft dysfunction. Am. J. Transplant. 2014, 14, 1152–1163. [Google Scholar] [CrossRef] [Green Version]
  83. Ichimura, T.; Asseldonk, E.J.; Humphreys, B.D.; Gunaratnam, L.; Duffield, J.S.; Bonventre, J.V. Kidney injury molecule-1 is a phosphatidylserine receptor that confers a phagocytic phenotype on epithelial cells. J. Clin. Invest. 2008, 118, 1657–1668. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Ichimura, T.; Bonventre, J.V.; Bailly, V.; Wei, H.; Hession, C.A.; Cate, R.L.; Sanicola, M. Kidney injury molecule-1 (KIM-1), a putative epithelial cell adhesion molecule containing a novel immunoglobulin domain, is up-regulated in renal cells after injury. J. Biol. Chem. 1998, 273, 4135–4142. [Google Scholar] [CrossRef] [Green Version]
  85. Humphreys, B.D.; Xu, F.; Sabbisetti, V.; Grgic, I.; Naini, S.M.; Wang, N.; Chen, G.; Xiao, S.; Patel, D.; Henderson, J.M.; et al. Chronic epithelial kidney injury molecule-1 expression causes murine kidney fibrosis. J. Clin. Invest. 2013, 123, 4023–4035. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Sabbisetti, V.S.; Waikar, S.S.; Antoine, D.J.; Smiles, A.; Wang, C.; Ravisankar, A.; Ito, K.; Sharma, S.; Ramadesikan, S.; Lee, M.; et al. Blood kidney injury molecule-1 is a biomarker of acute and chronic kidney injury and predicts progression to ESRD in type I diabetes. J. Am. Soc. Nephrol. 2014, 25, 2177–2186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Nowak, N.; Skupien, J.; Smiles, A.M.; Yamanouchi, M.; Niewczas, M.A.; Galecki, A.T.; Duffin, K.L.; Breyer, M.D.; Pullen, N.; Bonventre, J.V.; et al. Markers of early progressive renal decline in type 2 diabetes suggest different implications for etiological studies and prognostic tests development. Kidney Int. 2018, 93, 1198–1206. [Google Scholar] [CrossRef] [PubMed]
  88. Schrauben, S.J.; Shou, H.; Zhang, X.; Anderson, A.H.; Bonventre, J.V.; Chen, J.; Coca, S.; Furth, S.L.; Greenberg, J.H.; Gutierrez, O.M.; et al. Association of multiple plasma biomarker concentrations with progression of prevalent diabetic kidney disease: Findings from the Chronic Renal Insufficiency Cohort (CRIC) Study. J. Am. Soc. Nephrol. 2020, 32, 115–126. [Google Scholar] [CrossRef] [PubMed]
  89. Mori, Y.; Ajay, A.K.; Chang, J.H.; Mou, S.; Zhao, H.; Kishi, S.; Li, J.; Brooks, C.R.; Xiao, S.; Woo, H.M.; et al. KIM-1 mediates fatty acid uptake by renal tubular cells to promote progressive diabetic kidney disease. Cell Metab. 2021, 33, 1042–1061.e1047. [Google Scholar] [CrossRef] [PubMed]
  90. Zhao, X.; Chen, X.; Zhang, Y.; George, J.; Cobbs, A.; Wang, G.; Li, L.; Emmett, N. Kidney injury molecule-1 is upregulated in renal lipotoxicity and mediates palmitate-induced tubular cell injury and inflammatory response. Int. J. Mol. Sci. 2019, 20, 3406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Larrinaga, G.; Varona, A.; Pérez, I.; Sanz, B.; Ugalde, A.; Cándenas, M.L.; Pinto, F.M.; Gil, J.; López, J.I. Expression of cannabinoid receptors in human kidney. Histol. Histopathol. 2010, 25, 1133–1138. [Google Scholar] [CrossRef] [PubMed]
  92. Park, F.; Potukuchi, P.K.; Moradi, H.; Kovesdy, C.P. Cannabinoids and the kidney: Effects in health and disease. Am. J. Physiol. Renal Physiol. 2017, 313, F1124–F1132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Barutta, F.; Corbelli, A.; Mastrocola, R.; Gambino, R.; Di Marzo, V.; Pinach, S.; Rastaldi, M.P.; Perin, P.C.; Gruden, G. Cannabinoid receptor 1 blockade ameliorates albuminuria in experimental diabetic nephropathy. Diabetes 2010, 59, 1046–1054. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Jourdan, T.; Szanda, G.; Rosenberg, A.Z.; Tam, J.; Earley, B.J.; Godlewski, G.; Cinar, R.; Liu, Z.; Liu, J.; Ju, C.; et al. Overactive cannabinoid 1 receptor in podocytes drives type 2 diabetic nephropathy. Proc. Natl. Acad. Sci. USA. 2014, 111, E5420–E5428. [Google Scholar] [CrossRef] [Green Version]
  95. Lim, J.C.; Lim, S.K.; Han, H.J.; Park, S.H. Cannabinoid receptor 1 mediates palmitic acid-induced apoptosis via endoplasmic reticulum stress in human renal proximal tubular cells. J. Cell. Physiol. 2010, 225, 654–663. [Google Scholar] [CrossRef]
  96. Udi, S.; Hinden, L.; Earley, B.; Drori, A.; Reuveni, N.; Hadar, R.; Cinar, R.; Nemirovski, A.; Tam, J. Proximal Tubular Cannabinoid-1 Receptor Regulates Obesity-Induced CKD. J. Am. Soc. Nephrol. 2017, 28, 3518–3532. [Google Scholar] [CrossRef] [Green Version]
  97. Barutta, F.; Grimaldi, S.; Franco, I.; Bellini, S.; Gambino, R.; Pinach, S.; Corbelli, A.; Bruno, G.; Rastaldi, M.P.; Aveta, T.; et al. Deficiency of cannabinoid receptor of type 2 worsens renal functional and structural abnormalities in streptozotocin-induced diabetic mice. Kidney Int. 2014, 86, 979–990. [Google Scholar] [CrossRef] [Green Version]
  98. Barutta, F.; Piscitelli, F.; Pinach, S.; Bruno, G.; Gambino, R.; Rastaldi, M.P.; Salvidio, G.; Di Marzo, V.; Cavallo Perin, P.; Gruden, G. Protective role of cannabinoid receptor type 2 in a mouse model of diabetic nephropathy. Diabetes 2011, 60, 2386–2396. [Google Scholar] [CrossRef] [Green Version]
  99. Susztak, K.; Ciccone, E.; McCue, P.; Sharma, K.; Bottinger, E.P. Multiple metabolic hits converge on CD36 as novel mediator of tubular epithelial apoptosis in diabetic nephropathy. PLoS Med. 2005, 2, e45. [Google Scholar] [CrossRef] [Green Version]
  100. Susztak, K.; Böttinger, E.; Novetsky, A.; Liang, D.; Zhu, Y.Q.; Ciccone, E.; Wu, D.N.; Dunn, S.; McCue, P.; Sharma, K. Molecular profiling of diabetic mouse kidney reveals novel genes linked to glomerular disease. Diabetes 2004, 53, 784–794. [Google Scholar] [CrossRef] [Green Version]
  101. Wilson, P.C.; Wu, H.; Kirita, Y.; Uchimura, K.; Ledru, N.; Rennke, H.G.; Welling, P.A.; Waikar, S.S.; Humphreys, B.D. The single-cell transcriptomic landscape of early human diabetic nephropathy. Proc. Natl. Acad. Sci. USA. 2019, 116, 19619–19625. [Google Scholar] [CrossRef] [PubMed]
  102. Hunt, M.C.; Yamada, J.; Maltais, L.J.; Wright, M.W.; Podesta, E.J.; Alexson, S.E. A revised nomenclature for mammalian acyl-CoA thioesterases/hydrolases. J. Lipid Res. 2005, 46, 2029–2032. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Kirkby, B.; Roman, N.; Kobe, B.; Kellie, S.; Forwood, J.K. Functional and structural properties of mammalian acyl-coenzyme A thioesterases. Prog. Lipid Res. 2010, 49, 366–377. [Google Scholar] [CrossRef] [PubMed]
  104. Farese, R.V., Jr.; Walther, T.C. Lipid droplets finally get a little R-E-S-P-E-C-T. Cell 2009, 139, 855–860. [Google Scholar] [CrossRef] [Green Version]
  105. Xu, N.; Zhang, S.O.; Cole, R.A.; McKinney, S.A.; Guo, F.; Haas, J.T.; Bobba, S.; Farese, R.V., Jr.; Mak, H.Y. The FATP1-DGAT2 complex facilitates lipid droplet expansion at the ER-lipid droplet interface. J. Cell Biol. 2012, 198, 895–911. [Google Scholar] [CrossRef] [Green Version]
  106. Sztalryd, C.; Brasaemle, D.L. The perilipin family of lipid droplet proteins: Gatekeepers of intracellular lipolysis. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2017, 1862, 1221–1232. [Google Scholar] [CrossRef]
  107. Zechner, R.; Madeo, F.; Kratky, D. Cytosolic lipolysis and lipophagy: Two sides of the same coin. Nat. Rev. Mol. Cell Biol. 2017, 18, 671–684. [Google Scholar] [CrossRef]
  108. Urahama, Y.; Ohsaki, Y.; Fujita, Y.; Maruyama, S.; Yuzawa, Y.; Matsuo, S.; Fujimoto, T. Lipid droplet-associated proteins protect renal tubular cells from fatty acid-induced apoptosis. Am. J. Pathol. 2008, 173, 1286–1294. [Google Scholar] [CrossRef] [Green Version]
  109. Lubojemska, A.; Stefana, M.I.; Sorge, S.; Bailey, A.P.; Lampe, L.; Yoshimura, A.; Burrell, A.; Collinson, L.; Gould, A.P. Adipose triglyceride lipase protects renal cell endocytosis in a Drosophila dietary model of chronic kidney disease. PLoS Biol. 2021, 19, e3001230. [Google Scholar] [CrossRef]
  110. Pressly, J.D.; Gurumani, M.Z.; Varona Santos, J.T.; Fornoni, A.; Merscher, S.; Al-Ali, H. Adaptive and maladaptive roles of lipid droplets in health and disease. Am. J. Physiol. Cell Physiol. 2022, 322, C468–C481. [Google Scholar] [CrossRef]
  111. Garbarino, J.; Padamsee, M.; Wilcox, L.; Oelkers, P.M.; D’, A.D.; Ruggles, K.; Ramsey, N.; Jabado, O.; Turkish, A.; Sturley, S.L. Sterol and diacylglycerol acyltransferase deficiency triggers fatty acid mediated cell death. J. Biol. Chem. 2009, 284, 30994–31005. [Google Scholar] [CrossRef] [PubMed]
  112. Greenberg, A.S.; Coleman, R.A.; Kraemer, F.B.; McManaman, J.L.; Obin, M.S.; Puri, V.; Yan, Q.W.; Miyoshi, H.; Mashek, D.G. The role of lipid droplets in metabolic disease in rodents and humans. J. Clin. Invest. 2011, 121, 2102–2110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Olzmann, J.A.; Carvalho, P. Dynamics and functions of lipid droplets. Nat. Rev. Mol. Cell. Biol. 2019, 20, 137–155. [Google Scholar] [CrossRef] [PubMed]
  114. Valm, A.M.; Cohen, S.; Legant, W.R.; Melunis, J.; Hershberg, U.; Wait, E.; Cohen, A.R.; Davidson, M.W.; Betzig, E.; Lippincott-Schwartz, J. Applying systems-level spectral imaging and analysis to reveal the organelle interactome. Nature 2017, 546, 162–167. [Google Scholar] [CrossRef] [Green Version]
  115. Wang, H.; Sreenivasan, U.; Hu, H.; Saladino, A.; Polster, B.M.; Lund, L.M.; Gong, D.W.; Stanley, W.C.; Sztalryd, C. Perilipin 5, a lipid droplet-associated protein, provides physical and metabolic linkage to mitochondria. J. Lipid Res. 2011, 52, 2159–2168. [Google Scholar] [CrossRef] [Green Version]
  116. Benador, I.Y.; Veliova, M.; Mahdaviani, K.; Petcherski, A.; Wikstrom, J.D.; Assali, E.A.; Acin-Perez, R.; Shum, M.; Oliveira, M.F.; Cinti, S.; et al. Mitochondria bound to lipid droplets have unique bioenergetics, composition, and dynamics that support lipid droplet expansion. Cell Metab. 2018, 27, 869–885.e866. [Google Scholar] [CrossRef] [Green Version]
  117. Kien, B.; Kolleritsch, S.; Kunowska, N.; Heier, C.; Chalhoub, G.; Tilp, A.; Wolinski, H.; Stelzl, U.; Haemmerle, G. Lipid droplet-mitochondria coupling via perilipin 5 augments respiratory capacity but is dispensable for FA oxidation. J. Lipid Res. 2022, 63, 100172. [Google Scholar] [CrossRef]
  118. Chun, J.; Zhang, J.Y.; Wilkins, M.S.; Subramanian, B.; Riella, C.; Magraner, J.M.; Alper, S.L.; Friedman, D.J.; Pollak, M.R. Recruitment of APOL1 kidney disease risk variants to lipid droplets attenuates cell toxicity. Proc. Natl. Acad. Sci. U.S.A. 2019, 116, 3712–3721. [Google Scholar] [CrossRef] [Green Version]
  119. Chun, J.; Riella, C.; Chung, H.; Shah, S.; Wang, M.; Magraner, J.; Ribas, G.; Ribas, H.; Zhang, J.Y.; Alper, S.; et al. DGAT2 inhibition potentiates lipid droplet formation to reduce cytotoxicity in APOL1 kidney risk variants. J. Am. Soc. Nephrol. 2022, 33, 889–907. [Google Scholar] [CrossRef]
  120. Friedman, D.J.; Pollak, M.R. APOL1 nephropathy: From genetics to clinical applications. Clin. J. Am. Soc. Nephrol. 2021, 16, 294–303. [Google Scholar] [CrossRef]
  121. Keech, A.; Simes, R.J.; Barter, P.; Best, J.; Scott, R.; Taskinen, M.R.; Forder, P.; Pillai, A.; Davis, T.; Glasziou, P.; et al. Effects of long-term fenofibrate therapy on cardiovascular events in 9795 people with type 2 diabetes mellitus (the FIELD study): Randomised controlled trial. Lancet 2005, 366, 1849–1861. [Google Scholar] [CrossRef]
  122. Ginsberg, H.N.; Elam, M.B.; Lovato, L.C.; Crouse, J.R., III.; Leiter, L.A.; Linz, P.; Friedewald, W.T.; Buse, J.B.; Gerstein, H.C.; Probstfield, J.; et al. Effects of combination lipid therapy in type 2 diabetes mellitus. N. Engl. J. Med. 2010, 362, 1563–1574. [Google Scholar] [PubMed]
  123. Mychaleckyj, J.C.; Craven, T.; Nayak, U.; Buse, J.; Crouse, J.R.; Elam, M.; Kirchner, K.; Lorber, D.; Marcovina, S.; Sivitz, W.; et al. Reversibility of fenofibrate therapy-induced renal function impairment in ACCORD type 2 diabetic participants. Diabetes Care 2012, 35, 1008–1014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Forsblom, C.; Hiukka, A.; Leinonen, E.S.; Sundvall, J.; Groop, P.H.; Taskinen, M.R. Effects of long-term fenofibrate treatment on markers of renal function in type 2 diabetes: The FIELD Helsinki substudy. Diabetes Care. 2010, 33, 215–220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Kostapanos, M.S.; Florentin, M.; Elisaf, M.S. Fenofibrate and the kidney: An overview. Eur. J. Clin. Invest. 2013, 43, 522–531. [Google Scholar] [CrossRef] [PubMed]
  126. Frazier, R.; Mehta, R.; Cai, X.; Lee, J.; Napoli, S.; Craven, T.; Tuazon, J.; Safdi, A.; Scialla, J.; Susztak, K.; et al. Associations of fenofibrate therapy with incidence and progression of CKD in patients with type 2 diabetes. Kidney Int. Rep. 2019, 4, 94–102. [Google Scholar] [CrossRef] [Green Version]
  127. Baigent, C.; Landray, M.J.; Reith, C.; Emberson, J.; Wheeler, D.C.; Tomson, C.; Wanner, C.; Krane, V.; Cass, A.; Craig, J.; et al. The effects of lowering LDL cholesterol with simvastatin plus ezetimibe in patients with chronic kidney disease (Study of Heart and Renal Protection): A randomised placebo-controlled trial. Lancet 2011, 377, 2181–2192. [Google Scholar] [CrossRef] [Green Version]
  128. Herrington, W.G.; Emberson, J.; Mihaylova, B.; Blackwell, L.; Reith, C.; Solbu, M.D.; Mark, P.B.; Fellström, B.; Jardine, A.G.; Wanner, C.; et al. Impact of renal function on the effects of LDL cholesterol lowering with statin-based regimens: A meta-analysis of individual participant data from 28 randomised trials. Lancet Diabetes Endocrinol. 2016, 4, 829–839. [Google Scholar] [CrossRef] [Green Version]
  129. Shen, X.; Zhang, Z.; Zhang, X.; Zhao, J.; Zhou, X.; Xu, Q.; Shang, H.; Dong, J.; Liao, L. Efficacy of statins in patients with diabetic nephropathy: A meta-analysis of randomized controlled trials. Lipids Health Dis. 2016, 15, 179. [Google Scholar] [CrossRef] [Green Version]
  130. Qin, X.; Dong, H.; Fang, K.; Lu, F. The effect of statins on renal outcomes in patients with diabetic kidney disease: A systematic review and meta-analysis. Diabetes Metab. Res. Rev. 2017, 33, e2901. [Google Scholar] [CrossRef]
  131. Sandhu, S.; Wiebe, N.; Fried, L.F.; Tonelli, M. Statins for improving renal outcomes: A meta-analysis. J. Am. Soc. Nephrol. 2006, 17, 2006–2016. [Google Scholar] [CrossRef] [PubMed]
  132. Wanner, C.; Inzucchi, S.E.; Zinman, B. Empagliflozin and progression of kidney disease in type 2 diabetes. N. Engl. J. Med. 2016, 375, 1801–1802. [Google Scholar] [CrossRef] [PubMed]
  133. Perkovic, V.; Jardine, M.J.; Neal, B.; Bompoint, S.; Heerspink, H.J.L.; Charytan, D.M.; Edwards, R.; Agarwal, R.; Bakris, G.; Bull, S.; et al. Canagliflozin and renal outcomes in type 2 diabetes and nephropathy. N. Engl. J. Med. 2019, 380, 2295–2306. [Google Scholar] [CrossRef] [Green Version]
  134. Chertow, G.; Vart, P.; Jongs, N.; Toto, R.; Gorriz, J.L.; Hou, F.F.; McMurray, J.; Correa-Rotter, R.; Rossing, P.; Sjöström, C.D.; et al. Effects of dapagliflozin in stage 4 chronic kidney disease. J. Am. Soc. Nephrol. 2021, 32, 2352–2361. [Google Scholar] [CrossRef] [PubMed]
  135. Takata, T.; Isomoto, H. Pleiotropic effects of sodium-glucose cotransporter-2 inhibitors: Renoprotective mechanisms beyond glycemic control. Int. J. Mol. Sci. 2021, 22, 4374. [Google Scholar] [CrossRef] [PubMed]
  136. Wang, D.; Luo, Y.; Wang, X.; Orlicky, D.J.; Myakala, K.; Yang, P.; Levi, M. The sodium-glucose cotransporter 2 inhibitor dapagliflozin prevents renal and liver disease in Western diet induced obesity mice. Int. J. Mol. Sci. 2018, 19, 137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Hosokawa, K.; Takata, T.; Sugihara, T.; Matono, T.; Koda, M.; Kanda, T.; Taniguchi, S.; Ida, A.; Mae, Y.; Yamamoto, M.; et al. Ipragliflozin ameliorates endoplasmic reticulum stress and apoptosis through preventing ectopic lipid deposition in renal tubules. Int. J. Mol. Sci. 2019, 21, 190. [Google Scholar] [CrossRef] [Green Version]
  138. Shibusawa, R.; Yamada, E.; Okada, S.; Nakajima, Y.; Bastie, C.C.; Maeshima, A.; Kaira, K.; Yamada, M. Dapagliflozin rescues endoplasmic reticulum stress-mediated cell death. Sci. Rep. 2019, 9, 9887. [Google Scholar] [CrossRef] [Green Version]
  139. Takagi, S.; Li, J.; Takagaki, Y.; Kitada, M.; Nitta, K.; Takasu, T.; Kanasaki, K.; Koya, D. Ipragliflozin improves mitochondrial abnormalities in renal tubules induced by a high-fat diet. J. Diabetes Investig. 2018, 9, 1025–1032. [Google Scholar] [CrossRef] [Green Version]
  140. Wei, D.; Liao, L.; Wang, H.; Zhang, W.; Wang, T.; Xu, Z. Canagliflozin ameliorates obesity by improving mitochondrial function and fatty acid oxidation via PPARα in vivo and in vitro. Life Sci. 2020, 247, 117414. [Google Scholar] [CrossRef]
  141. Su, K.; Yi, B.; Yao, B.Q.; Xia, T.; Yang, Y.F.; Zhang, Z.H.; Chen, C. Liraglutide attenuates renal tubular ectopic lipid deposition in rats with diabetic nephropathy by inhibiting lipid synthesis and promoting lipolysis. Pharmacol. Res. 2020, 156, 104778. [Google Scholar] [CrossRef] [PubMed]
  142. Gagnon, L.; Leduc, M.; Thibodeau, J.F.; Zhang, M.Z.; Grouix, B.; Sarra-Bournet, F.; Gagnon, W.; Hince, K.; Tremblay, M.; Geerts, L.; et al. A newly discovered antifibrotic pathway regulated by two fatty acid receptors: GPR40 and GPR84. Am. J. Pathol. 2018, 188, 1132–1148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Li, Y.; Chung, S.; Li, Z.; Overstreet, J.M.; Gagnon, L.; Grouix, B.; Leduc, M.; Laurin, P.; Zhang, M.Z.; Harris, R.C. Fatty acid receptor modulator PBI-4050 inhibits kidney fibrosis and improves glycemic control. JCI insight 2018, 3, e120365. [Google Scholar] [CrossRef] [Green Version]
  144. Janiak, P.; Poirier, B.; Bidouard, J.P.; Cadrouvele, C.; Pierre, F.; Gouraud, L.; Barbosa, I.; Dedio, J.; Maffrand, J.P.; Le Fur, G.; et al. Blockade of cannabinoid CB1 receptors improves renal function, metabolic profile, and increased survival of obese Zucker rats. Kidney Int 2007, 72, 1345–1357. [Google Scholar] [CrossRef] [Green Version]
  145. Nam, D.H.; Lee, M.H.; Kim, J.E.; Song, H.K.; Kang, Y.S.; Lee, J.E.; Kim, H.W.; Cha, J.J.; Hyun, Y.Y.; Kim, S.H.; et al. Blockade of cannabinoid receptor 1 improves insulin resistance, lipid metabolism, and diabetic nephropathy in db/db mice. Endocrinology 2012, 153, 1387–1396. [Google Scholar] [CrossRef] [Green Version]
  146. Jenkin, K.A.; O’Keefe, L.; Simcocks, A.C.; Grinfeld, E.; Mathai, M.L.; McAinch, A.J.; Hryciw, D.H. Chronic administration of AM251 improves albuminuria and renal tubular structure in obese rats. J. Endocrinol. 2015, 225, 113–124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Wang, X.X.; Wang, D.; Luo, Y.; Myakala, K.; Dobrinskikh, E.; Rosenberg, A.Z.; Levi, J.; Kopp, J.B.; Field, A.; Hill, A.; et al. FXR/TGR5 dual agonist prevents progression of nephropathy in diabetes and obesity. J. Am. Soc. Nephrol. 2018, 29, 118–137. [Google Scholar] [CrossRef] [Green Version]
  148. Kim, Y.; Lim, J.H.; Kim, M.Y.; Kim, E.N.; Yoon, H.E.; Shin, S.J.; Choi, B.S.; Kim, Y.S.; Chang, Y.S.; Park, C.W. The adiponectin receptor agonist AdipoRon ameliorates diabetic nephropathy in a model of type 2 diabetes. J. Am. Soc. Nephrol. 2018, 29, 1108–1127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Lee, E.S.; Kang, J.S.; Kim, H.M.; Kim, S.J.; Kim, N.; Lee, J.O.; Kim, H.S.; Lee, E.Y.; Chung, C.H. Dehydrozingerone inhibits renal lipotoxicity in high-fat diet-induced obese mice. J. Cell. Mol. Med. 2021, 25, 8725–8733. [Google Scholar] [CrossRef] [PubMed]
  150. Gu, W.; Wang, X.; Zhao, H.; Geng, J.; Li, X.; Zheng, K.; Guan, Y.; Hou, X.; Wang, C.; Song, G. Resveratrol ameliorates diabetic kidney injury by reducing lipotoxicity and modulates expression of components of the junctional adhesion molecule-like/sirtuin 1 lipid metabolism pathway. Eur. J. Pharmacol. 2022, 918, 174776. [Google Scholar] [CrossRef] [PubMed]
  151. Nyambuya, T.M.; Nkambule, B.B.; Mazibuko-Mbeje, S.E.; Mxinwa, V.; Mokgalaboni, K.; Orlando, P.; Silvestri, S.; Louw, J.; Tiano, L.; Dludla, P.V. A meta-analysis of the impact of resveratrol supplementation on markers of renal function and blood pressure in type 2 diabetic patients on hypoglycemic therapy. Molecules 2020, 25, 5645. [Google Scholar] [CrossRef]
  152. Jao, T.M.; Nangaku, M.; Wu, C.H.; Sugahara, M.; Saito, H.; Maekawa, H.; Ishimoto, Y.; Aoe, M.; Inoue, T.; Tanaka, T.; et al. ATF6α downregulation of PPARα promotes lipotoxicity-induced tubulointerstitial fibrosis. Kidney Int. 2019, 95, 577–589. [Google Scholar] [CrossRef] [Green Version]
  153. Park, C.W.; Zhang, Y.; Zhang, X.; Wu, J.; Chen, L.; Cha, D.R.; Su, D.; Hwang, M.T.; Fan, X.; Davis, L.; et al. PPARα agonist fenofibrate improves diabetic nephropathy in db/db mice. Kidney Int. 2006, 69, 1511–1517. [Google Scholar] [CrossRef] [Green Version]
  154. Hong, Y.A.; Lim, J.H.; Kim, M.Y.; Kim, T.W.; Kim, Y.; Yang, K.S.; Park, H.S.; Choi, S.R.; Chung, S.; Kim, H.W.; et al. Fenofibrate improves renal lipotoxicity through activation of AMPK-PGC-1α in db/db mice. PLoS ONE 2014, 9, e96147. [Google Scholar] [CrossRef] [PubMed]
  155. Sohn, M.; Kim, K.; Uddin, M.J.; Lee, G.; Hwang, I.; Kang, H.; Kim, H.; Lee, J.H.; Ha, H. Delayed treatment with fenofibrate protects against high-fat diet-induced kidney injury in mice: The possible role of AMPK-autophagy. Am. J. Physiol. Renal Physiol. 2017, 312, F323–F334. [Google Scholar] [CrossRef] [Green Version]
  156. Aomura, D.; Harada, M.; Yamada, Y.; Nakajima, T.; Hashimoto, K.; Tanaka, N.; Kamijo, Y. Pemafibrate protects against fatty acid-induced nephropathy by maintaining renal fatty acid metabolism. Metabolites 2021, 11, 372. [Google Scholar] [CrossRef] [PubMed]
  157. Igweonu-Nwakile, E.O.; Ali, S.; Paul, S.; Yakkali, S.; Teresa Selvin, S.; Thomas, S.; Bikeyeva, V.; Abdullah, A.; Radivojevic, A.; Abu Jad, A.A.; et al. A systematic review on the safety and efficacy of PCSK9 inhibitors in lowering cardiovascular risks in patients with chronic kidney disease. Cureus 2022, 14, e29140. [Google Scholar] [CrossRef] [PubMed]
  158. Shroff, R.; Aitkenhead, H.; Costa, N.; Trivelli, A.; Litwin, M.; Picca, S.; Ali, A.; Peter, S.; Fatih, O.; Aleksandra, Z.; et al. Normal 25-hydroxyvitamin D levels are associated with less proteinuria and attenuate renal failure progression in children with CKD. J. Am. Soc. Nephrol. 2016, 27, 314–322. [Google Scholar] [CrossRef] [Green Version]
  159. Thomas, G.; Gadegbeku, C.A. Targeting the Standardized Blood Pressure: New Guidelines for Blood Pressure Management in Chronic Kidney Disease. Ann. Intern. Med. 2021, 174, 1321–1322. [Google Scholar] [CrossRef]
  160. Georgianos, P.I.; Roumeliotis, S.; Eleftheriadis, T.; Liakopoulos, V. Novel Therapeutic Strategies for Cardiorenal Protection in Patients with Type 2 Diabetes and Chronic Kidney Disease. Curr. Vasc. Pharmacol. 2022, 20, 117–120. [Google Scholar] [CrossRef]
  161. Wu, Y.; Lin, H.; Tao, Y.; Xu, Y.; Chen, J.; Jia, Y.; Zheng, Z. Network meta-analysis of mineralocorticoid receptor antagonists for diabetic kidney disease. Front. Pharmacol. 2022, 13, 967317. [Google Scholar] [CrossRef] [PubMed]
  162. Ishikawa, A.; Wada, T.; Nishimura, S.; Ito, T.; Okekawa, A.; Onogi, Y.; Watanabe, E.; Sameshima, A.; Tanaka, T.; Tsuneki, H.; et al. Estrogen regulates sex-specific localization of regulatory T cells in adipose tissue of obese female mice. PLoS ONE 2020, 15, e0230885. [Google Scholar] [CrossRef] [PubMed]
  163. Ferrere, G.; Leroux, A.; Wrzosek, L.; Puchois, V.; Gaudin, F.; Ciocan, D.; Renoud, M.L.; Naveau, S.; Perlemuter, G.; Cassard, A.M. Activation of kupffer cells is associated with a specific dysbiosis induced by fructose or high fat diet in mice. PLoS ONE 2016, 11, e0146177. [Google Scholar] [CrossRef] [Green Version]
  164. Boonyos, P.; Trakulnalueamsai, C.; Rungrod, A.; Chongthammakun, S.; Promdonkoy, B. Antagonistic effect of truncated fragments of Bacillus thuringiensis Vip3Aa on the larvicidal activity of its full-length protein. Protein Pept. Lett. 2021, 28, 131–139. [Google Scholar] [CrossRef]
  165. Liu, X.C.; Ducasa, G.M.; Mallela, S.K.; Kim, J.J.; Molina, J.; Mitrofanova, A.; Wilbon, S.S.; Ge, M.Y.; Fontanella, A.; Pedigo, C.; et al. Sterol-O-acyltransferase-1 has a role in kidney disease associated with diabetes and Alport syndrome. Kidney Int. 2020, 98, 1275–1285. [Google Scholar] [CrossRef]
  166. Ahowesso, C.; Black, P.N.; Saini, N.; Montefusco, D.; Chekal, J.; Malosh, C.; Lindsley, C.W.; Stauffer, S.R.; DiRusso, C.C. Chemical inhibition of fatty acid absorption and cellular uptake limits lipotoxic cell death. Biochem. Pharmacol. 2015, 98, 167–181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Lu, J.; Chen, P.P.; Zhang, J.X.; Li, X.Q.; Wang, G.H.; Yuan, B.Y.; Huang, S.J.; Liu, X.Q.; Jiang, T.T.; Wang, M.Y.; et al. GPR43 activation-mediated lipotoxicity contributes to podocyte injury in diabetic nephropathy by modulating the ERK/EGR1 pathway. Int. J. Biol. Sci. 2022, 18, 96–111. [Google Scholar] [CrossRef] [PubMed]
  168. Wang, X.X.; Edelstein, M.H.; Gafter, U.; Qiu, L.; Luo, Y.; Dobrinskikh, E.; Lucia, S.; Adorini, L.; D’Agati, V.D.; Levi, J.; et al. G protein-coupled bile acid receptor TGR5 activation inhibits kidney disease in obesity and diabetes. J. Am. Soc. Nephrol. 2016, 27, 1362–1378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Okada-Iwabu, M.; Yamauchi, T.; Iwabu, M.; Honma, T.; Hamagami, K.I.; Matsuda, K.; Yamaguchi, M.; Tanabe, H.; Kimura-Someya, T.; Shirouzu, M.; et al. A small-molecule AdipoR agonist for type 2 diabetes and short life in obesity. Nature. 2013, 503, 493–499. [Google Scholar] [CrossRef] [PubMed]
  170. Salami, M.; Salami, R.; Mafi, A.; Aarabi, M.H.; Vakili, O.; Asemi, Z. Therapeutic potential of resveratrol in diabetic nephropathy according to molecular signaling. Curr. Mol. Pharmacol. 2022, 15, 716–735. [Google Scholar] [CrossRef] [PubMed]
  171. Schelling, J.R. Tubular atrophy in the pathogenesis of chronic kidney disease progression. Pediatr. Nephrol. 2016, 31, 693–706. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Lipotoxicity in the podocyte. Under pathologic (albuminuric) circumstances, fatty acids and cholesterol esters are taken up by separate mechanisms. If the accumulated lipids exceed storage capacity within lipid droplets, toxic metabolites accumulate and lead to reactive oxygen species generation, endoplasmic reticulum stress, and ultimately podocyte cell death. Abbreviations: albumin (Alb), AMP-activated protein kinase (AMPK), ATP-binding cassette A1 (ABCA1), cannabinoid receptor type 1 (CB1), ceramide-1-phosphate (Cer1P), cholesterol ester (CE), endoplasmic reticulum (ER), fatty acid (FA), free fatty acid receptor-1 (FFA1), lipid droplet (LD), low density lipoprotein receptor (LDLR), reactive oxygen species (ROS), acid sphingomyelinase-like phosphodiesterase 3b (SMPDL3b).
Figure 1. Lipotoxicity in the podocyte. Under pathologic (albuminuric) circumstances, fatty acids and cholesterol esters are taken up by separate mechanisms. If the accumulated lipids exceed storage capacity within lipid droplets, toxic metabolites accumulate and lead to reactive oxygen species generation, endoplasmic reticulum stress, and ultimately podocyte cell death. Abbreviations: albumin (Alb), AMP-activated protein kinase (AMPK), ATP-binding cassette A1 (ABCA1), cannabinoid receptor type 1 (CB1), ceramide-1-phosphate (Cer1P), cholesterol ester (CE), endoplasmic reticulum (ER), fatty acid (FA), free fatty acid receptor-1 (FFA1), lipid droplet (LD), low density lipoprotein receptor (LDLR), reactive oxygen species (ROS), acid sphingomyelinase-like phosphodiesterase 3b (SMPDL3b).
Cells 11 03236 g001
Figure 2. Lipotoxicity in the proximal tubule cell. Under pathologic (albuminuric) circumustances, fatty acids are taken up by transporters at the basolateral and apical membranes. At the apical surface, fatty acids and albumin dissociate and are transported by separate mechanisms. If the accumulated lipids exceed the capacity to undergo β-oxidation within mitochondria or storage within lipid droplets, toxic fatty acid metabolites (ceramides, diacylglycerols, long-chain acyl-CoAs) cause apoptosis and tubular atrophy. Abbreviations: albumin (Alb), amnionless (Amn), cannabinoid receptor type 1 (CB1), ceramide (Cer), cubilin (Cub), diacylglycerols (DAG), fatty acid (FA), fatty acid transport protein (FATP), kidney injury molecule-1 (KIM-1), lipid droplet (LD), low density lipoprotein receptor (LDLR), long-chain acyl-CoAs (LC-CoA), megalin (Meg).
Figure 2. Lipotoxicity in the proximal tubule cell. Under pathologic (albuminuric) circumustances, fatty acids are taken up by transporters at the basolateral and apical membranes. At the apical surface, fatty acids and albumin dissociate and are transported by separate mechanisms. If the accumulated lipids exceed the capacity to undergo β-oxidation within mitochondria or storage within lipid droplets, toxic fatty acid metabolites (ceramides, diacylglycerols, long-chain acyl-CoAs) cause apoptosis and tubular atrophy. Abbreviations: albumin (Alb), amnionless (Amn), cannabinoid receptor type 1 (CB1), ceramide (Cer), cubilin (Cub), diacylglycerols (DAG), fatty acid (FA), fatty acid transport protein (FATP), kidney injury molecule-1 (KIM-1), lipid droplet (LD), low density lipoprotein receptor (LDLR), long-chain acyl-CoAs (LC-CoA), megalin (Meg).
Cells 11 03236 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Schelling, J.R. The Contribution of Lipotoxicity to Diabetic Kidney Disease. Cells 2022, 11, 3236. https://0-doi-org.brum.beds.ac.uk/10.3390/cells11203236

AMA Style

Schelling JR. The Contribution of Lipotoxicity to Diabetic Kidney Disease. Cells. 2022; 11(20):3236. https://0-doi-org.brum.beds.ac.uk/10.3390/cells11203236

Chicago/Turabian Style

Schelling, Jeffrey R. 2022. "The Contribution of Lipotoxicity to Diabetic Kidney Disease" Cells 11, no. 20: 3236. https://0-doi-org.brum.beds.ac.uk/10.3390/cells11203236

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