Next Article in Journal
A Compendium of AR Splice Variants in Metastatic Castration-Resistant Prostate Cancer
Next Article in Special Issue
Therapeutic Potential of Targeting Complement C5a Receptors in Diabetic Kidney Disease
Previous Article in Journal
CHIR99021-Treated Osteocytes with Wnt Activation in 3D-Printed Module Form an Osteogenic Microenvironment for Enhanced Osteogenesis and Vasculogenesis
Previous Article in Special Issue
Therapeutic Advances in Diabetic Kidney Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Perspective

The Role of Histone Modifications in the Pathogenesis of Diabetic Kidney Disease

by
Christodoula Kourtidou
and
Konstantinos Tziomalos
*
First Propedeutic Department of Internal Medicine, AHEPA Hospital, Medical School, Aristotle University of Thessaloniki, 54636 Thessaloniki, Greece
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(6), 6007; https://doi.org/10.3390/ijms24066007
Submission received: 17 February 2023 / Revised: 15 March 2023 / Accepted: 22 March 2023 / Published: 22 March 2023
(This article belongs to the Special Issue Advances in the Pathogenesis of Diabetic Kidney Disease)

Abstract

:
Diabetic kidney disease (DKD) is the leading cause of chronic kidney disease. The pathogenesis of DKD is multifactorial, with several molecular pathways implicated. Recent data suggest that histone modification plays an important role in the development and progression of DKD. Histone modification appears to induce oxidative stress, inflammation and fibrosis in the diabetic kidney. In the present review, we summarize the current knowledge on the association between histone modification and DKD.

1. Introduction

Chronic kidney disease (CKD) affects 9.1% of the population worldwide and is an important cause of morbidity and mortality [1]. Diabetic nephropathy is the leading cause of CKD and end-stage renal disease (ESRD) [2]. Even in prediabetes, there is evidence of kidney dysfunction [3]. The decline of estimated glomerular filtration rate (eGFR) and increased urinary albumin excretion are accompanied with an increased risk for progression to ESRD and death [4]. Moreover, CKD is a risk factor for cardiovascular disease (CVD) [5].
Epigenetic mechanisms include DNA methylation, histone modification, and microRNA (miRNA) [6]. Histone modification has been involved in the pathogenesis of diabetes mellitus but emerging data suggest that it also plays a role in the pathogenesis of DKD [7,8]. In the current review, we summarize the current evidence regarding the association between histone modifications and DKD. Of note, histone modification appears to be implicated in other microvascular complications of diabetes mellitus, i.e., neuropathy and retinopathy, and also in the pathogenesis of atherosclerosis in these patients [8], but these associations are outside of the scope of the present review.

2. Pathogenesis of DKD

The defining histopathologic features of DKD are glomerular basement membrane (GBM) thickening, mesangial matrix expansion, nodular glomerulosclerosis and tubular and vascular lesions [9]. p21 plays a key role in glomerular and mesangial cell hypertrophy [10]. The pathophysiology of DKD is characterized by impaired renal hemodynamics, oxidative stress, inflammation and a dysregulated renin–angiotensin–aldosterone system (RAAS) (Table 1 and Table 2) [11]. Glomerular hyperfiltration in diabetes predisposes to albuminuria, a decline in renal function and progressive kidney damage [12,13]. In addition to hyperglycemia, glomerular hyperfiltration in diabetic patients is also attributed to hyperaminoacidemia, hypertension and obesity [13]. Hyperinsulinemia stimulates endothelin 1 (ET-1) secretion in the endothelium and in turn increased ET-1 levels induce vasoconstriction which eventually leads to vascular dysfunction [14].
Intrarenal oxidative stress is another feature of DKD [15,35]. Oxidative stress predisposes to inflammation and renal fibrosis in DKD [15]. In diabetes, thioredoxin interacting protein (Txnip) is involved in the pathogenesis of oxidative stress, podocyte injury and apoptosis [16,17]. Mitochondrial injury in diabetes is present in renal tubular cells, podocytes, mesangial cells, and glomerular endothelial cells [36]. In turn, glomerular endothelial mitochondrial dysfunction might increase the risk for DKD [37]. Moreover, in the podocytes of DKD patients, mitochondrial dynamics are impaired [38].
Autophagy represents a mechanism of preservation of podocyte integrity [39]. Compromised autophagy is implicated in the pathogenesis of DKD [22,23]. Autophagy interacts with advanced glycation end-products (AGEs), oxidative stress, endoplasmic reticulum (ER) stress, RAAS activation and renal fibrosis in DKD [22,23]. mTORC1 negatively regulates autophagy by inhibiting the activity of the unc-51-like autophagy activating kinase 1 (ULK1) complex through direct phosphorylation [22]. Insulin also impairs autophagic activity in proximal tubule epithelial cells [24].
DKD is also characterized by the accumulation of extracellular matrix (ECM) proteins, collagen, laminin and fibronectin, in the glomerulus and the renal tubulointerstitium, which ultimately lead to glomerulosclerosis and tubulointerstitial fibrosis [40]. The pathogenesis of diabetic renal fibrosis involves several players, including transforming growth factor β (TGF-β), mitogen-activated protein kinase (MAPK), Wnt/β-catenin, phosphatidyl-inositol-3 kinase (PI3K)/Akt, JAK/STAT and Notch pathways, which also interact with each other [25]. Plasminogen activator inhibitor-1 (PAI-1) promotes ECM accumulation by regulating fibrinolysis and plasmin-mediated matrix metalloproteinase activation and is strongly induced in various forms of kidney diseases including DKD [26]. In patients with type 2 diabetes mellitus, PAI-1 levels were inversely linked to renal function [41].
Several immune cells of both the innate and adaptive immune system also contribute to renal injury in DKD [18,19]. In addition, cellular adhesion molecules including intracellular adhesion molecule 1 (ICAM-1), vascular cell adhesion molecule 1 (VCAM-1) and E-selectin, chemokines including monocyte chemoattractant protein-1 (MCP-1), and cytokines including interleukin-6 (IL-6) are inflammatory mediators of renal injury in DKD [19]. MCP-1 levels have been shown to predict renal disease progression in diabetic patients [42]. Nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB) is a transcriptional factor that regulates the expression of inflammatory cytokines, chemokines and cellular adhesion proteins that are involved in DKD [20]. The nucleotide binding and oligomerization domain-like receptor family pyrin domain-containing 3 (NLRP3) inflammasome, by interacting with the mitogen-activated protein kinase (MAPK), induces oxidative stress, NF-κB signaling pathways and inflammatory factors, thus resulting in impaired kidney functioning and structural changes that culminate in DKD development and progression [21].
Advanced glycation end-products (AGEs) induced by hyperglycemia play an important role in the pathogenesis of DKD [43,44]. AGEs appear to promote renal inflammation and fibrosis in DKD [44].
In patients with diabetes mellitus, the production of TGF-β is increased in the kidney, which then facilitates the development of early manifestations of DKD [45]. TGF-β was shown to be upregulated in the glomerular and tubulointerstitial compartments of diabetic animal models [46]. Moreover, TGF-β is involved in the pathogenesis of renal hypertrophy, glomerulosclerosis and tubulointerstitial fibrosis observed in DKD [47]. The TGF-β/Smad signaling plays a crucial role in renal fibrosis, inflammation [48,49] and podocyte apoptosis [50].
Epithelial-to-mesenchymal transition (EMT), which occurs in tubular cells after injury, refers to a change in phenotype that is characterized by the conversion of epithelial markers to mesenchymal features and generates kidney fibrosis [51]. The downregulation of e-cadherin is observed in EMT [52]. It has been reported that endothelial-to-mesenchymal transition (EndMT) in glomeruli, a subtype of EMT, might be involved in the pathogenesis of DKD [53,54].

3. Histone Modification and DKD

Histone modifications regulate gene transcription, DNA repair and DNA replication [55,56]. Post-translational modifications (PTMs) occur at the residues of N-terminal tails of the histone proteins while modifications are also detected in the core regions of the histones [57]. The major histone PTMs are acetylation, methylation, ubiquitination and phosphorylation [58].
There is accumulating evidence that PTMs are involved in the regulation of genes associated with the pathogenesis of diabetes [7,8]. The PTMs of several genes related to diabetes and inflammation have been identified in monocytes and lymphocytes under diabetic conditions [59,60]. The persistent expression of genes and phenotypes created by former exposure to hyperglycemia, despite the subsequent achievement of glycemic control, a phenomenon termed metabolic memory, might be caused by epigenetic mechanisms and be associated with DKD [61,62]. Several studies demonstrated the non-histone modifications of histone-modifying enzymes in the setting of DKD [63,64,65].

4. Histone Acetylation and DKD (Table 3)

Histone acetylation can unravel chromatin and enable the binding of transcription factors and cofactors that lead to gene expression [55]. Histone acetylation, such as H3K9Ac, H3K14Ac and H4KAc, is generally correlated with permissive gene expression, while histone deacetylation is mainly associated with gene transcriptional repression [66]. Histone acetylation is catalyzed by histone acetyltransferases (HATs) and histone deacetylation by histone deacetylases (HDACs) [55]. p300/CREB-binding protein (CBP) and p300/CBP-associated factor (PCAF) are some of the known HATs [67]. Among the four classes of HDACs, the sirtuins (Sirt 1–7) belong in the third class [67].
Studies in db/db diabetic mice reported that early glomerulosclerosis was associated with reduced histone H3K9 and H3K23 acetylation [68]. Other studies showed that acetylation levels of H3K9, H3K18 and H3K23 were increased in the kidneys of diabetic mice [29,69]. Moreover, H3K27ac levels in diabetic mice are increased compared with non-diabetic mice [70]. Moreover, the increased H3K9/14Ac levels in the diabetic kidney mainly involved the nuclei of both glomerular and renal tubular cells [71]. Others indicated that hyperglycemia induced histone acetylation at the H3K18 and H3K9/14 sites in mesangial cells [72].
Table 3. Effects of histone acetylation on the pathogenesis of diabetic kidney disease.
Table 3. Effects of histone acetylation on the pathogenesis of diabetic kidney disease.
Histone AcetylationGene/Molecule TargetImplicated MechanismHistone Acetyltranserases/Deacetylases or Their Inhibitors
H3K9/14AcPromoters of PAI-1 and p21 genes in mesangial cellsECM accumulation and hypertrophyp300/CBP [31]
Fibrinonectin protein expression in mesangial cellsRenal fibrosisLys-CoA [72]
BMP-7 gene promoter in tubular cellsRenal fibrosisSFN [27]
H3K18AcMCP-1, ICAM-1, VCAM-1 and iNOS in mesangial cellsMesangial cell inflammationApelin [29]
H3K9AcMCP-1, ICAM-1, VCAM-1 and iNOS in mesangial cellsMesangial cell inflammationApelin [29]
Promoters of Notch1 and Notch4 in hyperglycemia-treated podocytesApoptosis and inflammation of podocytesSirt6 [33]
Timp1 promoter in proximal tubular cells
miR-29a proximal promoter
Tubular fibrosisSirt6 [73]
Podocyte apoptosisHDAC4 [34]
PAI-1: Plasminogen activator inhibitor-1, ECM: extracellular matrix, p300/CBP: p300/CREB-binding protein, BMP-7: bone morphologic protein 7, SFN: sulforaphane, MCP-1: monocyte chemoattractant protein-1, ICAM-1: intracellular adhesion molecule 1, VCAM-1: vascular cell adhesion molecule 1, iNOS: inducible nitric oxide synthases, Sirt6: sirtuin 6.

5. Histone Acetylation and Renal Inflammation in DKD

In the context of hyperglycemia, increased H3K9 and H3K18 acetylation levels in the mesangial cells were associated with the production of inflammatory mediators including MCP-1, ICAM-1, VCAM-1 and inducible nitric oxide synthases (iNOS) [29]. Furthermore, ChIP assays revealed that PAI-induced enrichment of H3K18ac at the promoters of ICAM-1 and MCP-1 was mitigated after PCAF knockdown along with altered transcription levels [69]. In vitro studies demonstrated elevated levels of H3K9ac in the promoters of Notch1 and Notch4 in hyperglycemia-treated podocytes, whereas Sirt6 suppressed Notch1 and Notch4 transcription by deacetylating histone H3K9 and protected podocytes from apoptosis and inflammation through inhibition of the Notch pathway [33].

6. Histone Acetylation and Renal Fibrosis in DKD

The enrichment of H3K9/14Ac at the promoters of PAI-1 and p21 genes in rat mesangial cells exposed to hyperglycemia and glomeruli from diabetic mice were associated with the increased expression of these genes [31]. In addition, the elevated H3K9/14Ac levels on the gene promoters of connective tissue growth factor (CTGF), PAI-1 and fibronectin 1 (FN-1) in the kidneys of diabetic mice were associated with activation of HAT p300/CBP [71]. ChIP assays found elevated levels of acetylated H3 at the promoter sites of the ET-1, FN genes in human umbilical vein endothelial cells (HUVECs) [74]. Under hyperglycemic and hyperinsulinemic conditions in kidneys the, levels of histone H3 acetylation in the fibrillin 1 gene were increased [75]. In another study in renal tubular cells, exposure to hyperglycemia augmented the enrichment of H3K18Ac, H3K27Ac and H3K4Me3 but not H3K9Ac on the collagen gene promoters [76]. Claudin-1 expression was regulated through the deacetylation of histone H3 and H4 by Sirt1 while Claudin-1 expression in podocytes induced podocyte effacement and resulted in albuminuria [77]. Sirt6 increased H3K9 acetylation in the promoter region of the pro-fibrotic factor Timp1 leading to the upregulation of Timp1 that may inactivate matrix metalloproteinases (MMPs) and reduced fibrous tissue degradation [73]. In another study, 12(S)-hydroxyeicosatetraenoic acid (12(S)-HETE), an oxidized lipid which promotes oxidative stress, induced profibrotic gene expression in association with the augmented enrichment of their promoter by H3K9Ac and H3K4me in rat mesangial cell [78]. In mesangial cells, the inhibition of H3K9/14 acetylation by Lys-CoA, an inhibitor of CBP/p300 HAT activity, led to diminished FN protein expression [72]. In vitro, trichostatin A (TSA)-induced acetylation of H3 was correlated with the overexpression of E-cadherin and decreased expression of the FN gene in TGF-β1-treated tubular epithelial cells [79]. In addition, TSA treatment increased the acetylation of histones H3 and H4 in the E-cadherin promoter, H3 in the Id2 promoter and H4 in the bone morphologic protein 7 (BMP-7) promoter and subsequently regulated the expression of Id2 and BMP-7 [80]. In another study, sulforaphane (SFN) treatment counteracted the diabetes-induced decline in H3K9/14Ac expression and H3K9/14Ac levels in the BMP-7 gene promoter, which was accompanied by BMP-7 upregulation [27].

7. Histone Acetylation and Apoptosis in DKD

Blockade of HDAC4 signaling increased the H3K9Ac levels in the miR-29a proximal promoter and miR-29a transcription in hyperglycemia-treated podocytes and attenuated the hyperglycemia-induced apoptosis [34]. In another study, the acetylation of histone H4 in the glucose-regulated protein (GRP78) promoter, was increased in the DKD group, compared with controls; valproate reversed apoptosis by increasing the acetylation of histone H4 in the GRP78 promoter and decreasing the acetylation of histone H4 in the C/EBP-homologous protein (CHOP) promoter, two promoters that encode proteins related to ER stress [81].

8. Histone Acetylation and Other Manifestations of DKD

Hyperglycemia-induced Txnip expression was associated with the enrichment of H3K9ac at the promoter region of the gene in the kidneys of diabetic mice [82]. ChIP assays showed increased levels of H3K18Ac and H2BK5Ac in the membrane metalloendopeptidase (Mme) promoter region in the kidneys of diabetic rats but enrichment of H2AK5Ac remained unchanged on Mme promoter in kidneys [83]. The expression of ULK1 in diabetic mice was decreased compared with that in non-diabetic mice, whereas ChIP assays also revealed reduced levels of H3K27Ac on of the ULK1 promoter sequences in diabetic mice [84].

9. Histone Methylation and DKD (Table 4)

Histone methylation takes place in arginines, lysines and histidines on histone tails [85]. Histone methylation modifications promote or repress gene transcription [86]. Lysine methylation is catalyzed by methyltransferases (KMTs) and is reversed by demethylases (KDMs) [87,88,89]. SET domain-containing enzymes are one class of KMTs [87].
KDM6A expression was increased and H3K27me2/3 levels were decreased in the kidneys of diabetic mice [90] and in hyperglycemia-treated podocytes and renal tubular cells [91,92]. In addition, in kidney tissues from patients with DKD, there was a reduction of podocyte, glomerular cell and tubular cell H3K27me3 [93]. Elevated H3K4me1 and H3K4me3 levels and decreases in H3K27me3 levels were detected in the kidneys of diabetic animals compared with non-diabetic controls [94].
Table 4. Effects of histone methylation on the pathogenesis of diabetic kidney disease.
Table 4. Effects of histone methylation on the pathogenesis of diabetic kidney disease.
Histone MethylationGene/Molecule TargetImplicated MechanismHistone Methyltranserases/Demethylases or Their Inhibitors
H3K9me2/3p21 gene promoter in rat mesangial cellsMesangial cell hypertrophySET7/9 [32]
H3K4me1/2/3p21 gene promoter
in rat mesangial cells
Mesangial cell hypertrophySET7/9 [32]
H3K4me3Txnip promoter in mesangial cellsRenal inflammationC646 [95]
Spp1 gene hyperglycemia-treated mesangial cellsMM-102 [94]
H3K4me1MCP-1 promoter in db/db miceRenal inflammationSET7/9 [30]
Txnip promoter in mesangial cellsRenal inflammationC646 [95]
H3K27me3Promoters of IL1b and IL6 in renal mesangial cellsRenal inflammationPalmitic acid, KDM6A [90]
FN and PAI-1 expressionMesangial cell hypertrophyKMT6A [96]
HES1 promoter hyperglycemia-treated mesangial cellsMesangial cell proliferation and fibrosisKMT6A [97]
α-SMA, collagen I, and FN in rat renal interstitial fibroblast cellsRenal fibrosisKMT6A [28]
Pax6, TxnIP promoters in renal podocytesPodocyte oxidative stressKMT6A [98]
MCP-1: monocyte chemoattractant protein-1, Spp1: secreted phosphoprotein 1, IL1b: interleukin 1b, IL6: interleukin 6, KDM6A: lysine demethylase 6A, FN: fibronectin, PAI-1: plasminogen activator inhibitor-1, KMT6A: histone-lysine N-methyltrasferase, HES1: enhancer of split 1, α-SMA: smooth muscle alpha-actin, Pax6: paired box 6.

10. Histone Methylation and Renal Inflammation in DKD

Palmitic acid (PA) stimulation decreased the enrichment of H3K27me3 on the promoters of IL1b and IL6, whereas the overexpression of KDM6A further decreased H3K27me3 levels on their promoters in renal mesangial cells [90]. In contrast, under PA treatment, KDM6A knockdown increased the enrichment of H3K27me3 on the promoters of IL1b and IL6 [90]. In vitro studies in renal tubular cells exposed to hyperglycemia suggest that the secretion of IL6 and MCP-1 increases and inflammation stimulated the expression of KMT1A which catalyzed H3K9me3 to inhibit the transcription of inflammatory genes [99]. In another study, SET domain-containing protein 8 (SETD8) overexpression reduced hyperglycemia-induced IL-1β and IL-18 expression by attenuating hyperglycemia-induced MARK4 expression and inactivating the NLRP3 inflammasome, whereas H4K20me1, a downstream target of SETD8, was enriched in the MARK4 promoter region in HUVECs [100]. Furthermore, ChIP assays demonstrated that H4K20me1 levels were increased at the promoter region of E26 transformation-specific sequence transcription factor-1 (ESE-1), in HUVECs exposed to hyperglycemia [101]. Set7-induced enrichment of H3K4m1 on NF-kB p65 gene promoter led to the upregulation of ICAM-1, MCP-1, iNOS and cyclooxygenase-2 (COX-2) in peripheral blood monocytes [102]. Others reported that SET7/9 recruitment and H3K4me1 levels at MCP-1 promoters in diabetic mice were markedly higher compared with non-diabetic animals; the ER stress resulted in the enrichment of H3K4me1 at MCP-1 promoter in the kidneys of diabetic db/db mice through the preferential induction of SET7/9 [30]. Finally, renal hypertrophy in diabetes was linked to the upregulation of profibrotic and proinflammatory genes and reduced levels of H3K27me3 and KMT6A in these gene promoters [103].

11. Histone Methylation and Renal Fibrosis in DKD

The levels of H4K20me1, a downstream target of KMT5A, decreased in hyperglycemia-treated HUVECs and the levels of H4K20me1 were enriched in the promoter region of enolase 1 (ENO1), regulating the ENO1 levels and thereby inducing EndMT [104]. On the other hand, the expression of e-cadherin was decreased and its promoter methylation was upregulated with increasing KDM6A expression, but there was no relationship with the level of H3K27me2/3 in the E-cadherin promoter [92]. Hyperglycemia decreased KMT1A and histone H3K9me3 levels in the FN and p21 gene promoters [105]. On the other hand, the augmented expression of KMT1A increased the H3K9me3 levels in the FN and p21 promoters and protected against hyperglycemia-induced cell hypertrophy [105]. Hyperglycemia increased levels of KMT6A, resulting in the enrichment of H3K27me3 at the deptor promoter leading to decreased deptor expression, which led to the activation of mTORC1 and mTORC2, controlled mesangial cell hypertrophy and FN and PAI-1 expression [96]. In accordance with these findings, KMT6A inhibited mesangial cell proliferation and fibrosis-related protein expression in hyperglycemia-treated mesangial cells by downregulating hairy and enhancer of split 1 (HES1) expression; a ChiP assay showed that the KMT6A and H3K27me3 levels were increased in the HES1 promoter and enrichment was intensified by KMT6A overexpression [97]. Moreover, EZH2 knockdown significantly decreased the levels of H3K27me3 and downregulated the levels of smooth muscle alpha-actin (α-SMA), collagen I, and FN in rat renal interstitial fibroblast cells, whereas the silencing of KMT6A repressed the TGFβ1-stimulated expressions of α-SMA, collagen I, and FN [28]. In animal models of type 1 diabetes mellitus, the levels of H3K27m3 were reduced in MCP-1, vimentin and the fibrosis marker Fsp1 genes, while the levels of H3K4m2 were increased [106]. More recently, it was reported that histone H3 K79 methyltransferase Dot1l has an antifibrotic effect by repressing Edn1, which encodes endothelin-1 [95]. Moreover, Dot1l interacts with HDAC2 to regulate Edn1 transcription [95]. Notably, HDAC2 was also shown to protect against renal ischemia/reperfusion injury by suppressing Edn1 transcription [98].

12. Histone Methylation and Oxidative Stress in DKD

In the kidneys of diabetic mice, hyperglycemia-induced Txnip expression was associated with the stimulation of activating histone marks H3K4me3 and H3K4me1, as well as with a decrease in the repressive histone mark H3K27me3 at the promoter of the gene [95]. In another study, KMT6A downregulated TxnIP expression through H3K27me3 enrichment in the promoter region of the transcription factor paired box 6 (Pax6), whereas the depletion of KMT6A led to an increase in the hyperglycemia-induced reactive oxygen species (ROS) levels and cellular death [98]. In addition, the depletion or inhibition of KMT6A attenuated the increase of ROS in hyperglycemia-treated human renal tubular epithelial cells, while the KMT6A knockdown attenuated the reduced expression of silent information regulator 1 (SIRT1) by hyperglycemia treatment [107]. The restoration of SIRT1 expression was in line with the elimination of H3K27me3 from the SIRT1 promoter and the inhibition of intracellular ROS levels [107].

13. Histone Methylation and Glomerular and Mesangial Hypertrophy in DKD

TGF-β1 stimulated the expression of the p21 gene in rat mesangial cells, while the levels of H3K9me2/3 and H3K4me1/2/3 at p21 gene promoter were correlated with p21 gene expression [32]. Similarly, the reduced H3K9me2 level in the p21 gene promoter in the glomeruli of type 1 diabetic rats was inversely correlated with p21 gene upregulation [108]. The increased H3K4me3 level in secreted phosphoprotein 1 (Spp1), the osteopontin gene, promoter were associated with Spp1 gene enhanced expression in hyperglycemia-treated human mesangial cells, whereas the levels of histone marks were linked to increases of Spp1 expression in animal studies [94]. Diminished podocyte H3K27me3 was correlated with glomerular injury while H3K27me3 controlled Notch signaling and podocyte dedifferentiation [93].

14. Histone Methylation and Other Manifestations of DKD

SIRT-6 induction of activated protein kinase (AMPK) signaling, through H3K9 and H3K56, contributed to hyperglycemia-induced mitochondrial dysfunction and apoptosis in podocytes [109]. In another study, ChIP assays demonstrated the enrichment of H3K4me1 levels at the MAP4K3 gene and the upregulation of MAP4K3 in diabetic mice [110]. MAP4K3 knockdown decreased HYPERGLYCEMIA-induced apoptosis while MAP4K3 downregulation counteracted the expression of apoptosis markers [110]. Histone H2AK119 and H2BK120 ubiquitination regulated H3K4Me2 and H3K9Me2 chromatin marks by modifying the expression of their respective HMTs, SET7/9 and KMT1A [111]. The increased expression of SET7/9 and decreased expression of KMT1A could be associated with the increased enrichment of H3K4Me2, and decreased H3K9Me2 level at Collagen Type I Alpha 1 Chain (Col1a1) gene promoter and increased Col1a1 gene expression in the diabetic kidneys [111].

15. Conclusions

Histone modification appears to play an important role in the pathogenesis of DKD. Indeed, histone acetylation and methylation are implicated in the propagation of inflammation, oxidative stress and fibrosis in the kidney. Even though the exact mechanism(s) through which impaired epigenetic landscape contributes to DKD is still unclear, it appears that an altered enhancer landscape plays an important role [112,113,114]. It remains to be established whether targeting this pathway might prevent the development or delay the progression of DKD. Of note, it is well established that DKD is associated with an increased risk for CVD [115]. In this context, somatic mutations, particularly in DNA methyltransferase 3 alpha and TET methylcytosine dioxygenase 2, are also implicated in the pathogenesis of both atherosclerosis and DKD, partly by a pro-inflammatory effect [116,117].

Author Contributions

Writing—original draft preparation, C.K.; writing—review and editing, K.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created in this paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. GBD Chronic Kidney Disease Collaboration. Global, regional, and national burden of chronic kidney disease, 1990–2017: A systematic analysis for the Global Burden of Disease Study 2017. Lancet 2020, 395, 709–733. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Reutens, A.T. Epidemiology of diabetic kidney disease. Med. Clin. N. Am. 2013, 97, 1–18. [Google Scholar] [CrossRef] [PubMed]
  3. Sasson, A.N.; Cherney, D.Z. Renal hyperfiltration related to diabetes mellitus and obesity in human disease. World J. Diabetes 2012, 3, 1–6. [Google Scholar] [CrossRef] [PubMed]
  4. Astor, B.C.; Matsushita, K.; Gansevoort, R.T.; van der Velde, M.; Woodward, M.; Levey, A.S.; Jong, P.E.; Coresh, J.; Chronic Kidney Disease Prognosis Consortium. Lower estimated glomerular filtration rate and higher albuminuria are associated with mortality and end-stage renal disease. A collaborative meta-analysis of kidney disease population cohorts. Kidney Int. 2011, 79, 1331–1340. [Google Scholar] [CrossRef] [Green Version]
  5. Gansevoort, R.T.; Correa-Rotter, R.; Hemmelgarn, B.R.; Jafar, T.H.; Heerspink, H.J.; Mann, J.F.; Matsushita, K.; Wen, C.P. Chronic kidney disease and cardiovascular risk: Epidemiology, mechanisms, and prevention. Lancet 2013, 382, 339–352. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, L.; Lu, Q.; Chang, C. Epigenetics in Health and Disease. Adv. Exp. Med. Biol. 2020, 1253, 3–55. [Google Scholar] [CrossRef]
  7. Zhang, H.; Pollin, T.I. Epigenetics Variation and Pathogenesis in Diabetes. Curr. Diab. Rep. 2018, 18, 121. [Google Scholar] [CrossRef] [PubMed]
  8. Reddy, M.A.; Zhang, E.; Natarajan, R. Epigenetic mechanisms in diabetic complications and metabolic memory. Diabetologia 2015, 58, 443–455. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Tervaert, T.W.; Mooyaart, A.L.; Amann, K.; Cohen, A.H.; Cook, H.T.; Drachenberg, C.B.; Ferrario, F.; Fogo, A.B.; Haas, M.; de Heer, E.; et al. Pathologic classification of diabetic nephropathy. J. Am. Soc. Nephrol. 2010, 21, 556–563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Fan, Y.P.; Weiss, R.H. Exogenous attenuation of p21(Waf1/Cip1) decreases mesangial cell hypertrophy as a result of hyperglycemia and IGF-1. J. Am. Soc. Nephrol. 2004, 15, 575–584. [Google Scholar] [CrossRef] [Green Version]
  11. Lin, Y.C.; Chang, Y.H.; Yang, S.Y.; Wu, K.D.; Chu, T.S. Update of pathophysiology and management of diabetic kidney disease. J. Formos. Med. Assoc. 2018, 117, 662–675. [Google Scholar] [CrossRef] [PubMed]
  12. Tonneijck, L.; Muskiet, M.H.; Smits, M.M.; van Bommel, E.J.; Heerspink, H.J.; van Raalte, D.H.; Joles, J.A. Glomerular Hyperfiltration in Diabetes: Mechanisms, Clinical Significance, and Treatment. J. Am. Soc. Nephrol. 2017, 28, 1023–1039. [Google Scholar] [CrossRef] [Green Version]
  13. Alicic, R.Z.; Rooney, M.T.; Tuttle, K.R. Diabetic Kidney Disease: Challenges, Progress, and Possibilities. Clin. J. Am. Soc. Nephrol. 2017, 12, 2032–2045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Potenza, M.A.; Gagliardi, S.; Nacci, C.; Carratu, M.R.; Montagnani, M. Endothelial dysfunction in diabetes: From mechanisms to therapeutic targets. Curr. Med. Chem. 2009, 16, 94–112. [Google Scholar] [CrossRef]
  15. Jha, J.C.; Banal, C.; Chow, B.S.; Cooper, M.E.; Jandeleit-Dahm, K. Diabetes and Kidney Disease: Role of Oxidative Stress. Antioxid. Redox Signal. 2016, 25, 657–684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Gao, P.; He, F.F.; Tang, H.; Lei, C.T.; Chen, S.; Meng, X.F.; Su, H.; Zhang, C. NADPH oxidase-induced NALP3 inflammasome activation is driven by thioredoxin-interacting protein which contributes to podocyte injury in hyperglycemia. J. Diabetes Res. 2015, 2015, 504761. [Google Scholar] [CrossRef] [Green Version]
  17. Shah, A.; Xia, L.; Masson, E.A.; Gui, C.; Momen, A.; Shikatani, E.A.; Husain, M.; Quaggin, S.; John, R.; Fantus, I.G. Thioredoxin-Interacting Protein Deficiency Protects against Diabetic Nephropathy. J. Am. Soc. Nephrol. 2015, 26, 2963–2977. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Zheng, Z.; Zheng, F. Immune Cells and Inflammation in Diabetic Nephropathy. J. Diabetes Res. 2016, 2016, 1841690. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Rayego-Mateos, S.; Morgado-Pascual, J.L.; Opazo-Ríos, L.; Guerrero-Hue, M.; García-Caballero, C.; Vázquez-Carballo, C.; Mas, S.; Sanz, A.B.; Herencia, C.; Mezzano, S.; et al. Pathogenic Pathways and Therapeutic Approaches Targeting Inflammation in Diabetic Nephropathy. Int. J. Mol. Sci. 2020, 21, 3798. [Google Scholar] [CrossRef]
  20. Navarro-González, J.F.; Mora-Fernández, C.; Muros de Fuentes, M.; García-Pérez, J. Inflammatory molecules and pathways in the pathogenesis of diabetic nephropathy. Nat. Rev. Nephrol. 2011, 7, 327–340. [Google Scholar] [CrossRef]
  21. Qiu, Y.Y.; Tang, L.Q. Roles of the NLRP3 inflammasome in the pathogenesis of diabetic nephropathy. Pharmacol. Res. 2016, 114, 251–264. [Google Scholar] [CrossRef] [PubMed]
  22. Ding, Y.; Choi, M.E. Autophagy in diabetic nephropathy. J. Endocrinol. 2015, 224, R15–R30. [Google Scholar] [CrossRef] [Green Version]
  23. Yang, D.; Livingston, M.J.; Liu, Z.; Dong, G.; Zhang, M.; Chen, J.K.; Dong, Z. Autophagy in diabetic kidney disease: Regulation, pathological role and therapeutic potential. Cell. Mol. Life Sci. 2018, 75, 669–688. [Google Scholar] [CrossRef] [PubMed]
  24. Sakai, S.; Yamamoto, T.; Takabatake, Y.; Takahashi, A.; Namba-Hamano, T.; Minami, S.; Fujimura, R.; Yonishi, H.; Matsuda, J.; Hesaka, A.; et al. Proximal Tubule Autophagy Differs in Type 1 and 2 Diabetes. J. Am. Soc. Nephrol. 2019, 30, 929–945. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, Y.; Jin, D.; Kang, X.; Zhou, R.; Sun, Y.; Lian, F.; Tong, X. Signaling Pathways Involved in Diabetic Renal Fibrosis. Front. Cell Dev. Biol. 2021, 9, 696542. [Google Scholar] [CrossRef]
  26. Nicholas, S.B.; Aguiniga, E.; Ren, Y.; Kim, J.; Wong, J.; Govindarajan, N.; Noda, M.; Wang, W.; Kawano, Y.; Collins, A.; et al. Plasminogen activator inhibitor-1 deficiency retards diabetic nephropathy. Kidney Int. 2005, 67, 1297–1307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Kong, L.; Wang, H.; Li, C.; Cheng, H.; Cui, Y.; Liu, L.; Zhao, Y. Sulforaphane Ameliorates Diabetes-Induced Renal Fibrosis through Epigenetic Up-Regulation of BMP-7. Diabetes Metab. J. 2021, 45, 909–920. [Google Scholar] [CrossRef]
  28. Zhou, X.; Zang, X.; Ponnusamy, M.; Masucci, M.V.; Tolbert, E.; Gong, R.; Zhao, T.C.; Liu, N.; Bayliss, G.; Dworkin, L.D.; et al. Enhancer of Zeste Homolog 2 Inhibition Attenuates Renal Fibrosis by Maintaining Smad7 and Phosphatase and Tensin Homolog Expression. J. Am. Soc. Nephrol. 2016, 27, 2092–2108. [Google Scholar] [CrossRef] [Green Version]
  29. Chen, H.; Li, J.; Jiao, L.; Petersen, R.B.; Li, J.; Peng, A.; Zheng, L.; Huang, K. Apelin inhibits the development of diabetic nephropathy by regulating histone acetylation in Akita mouse. J. Physiol. 2014, 592, 505–521. [Google Scholar] [CrossRef] [Green Version]
  30. Chen, J.; Guo, Y.; Zeng, W.; Huang, L.; Pang, Q.; Nie, L.; Mu, J.; Yuan, F.; Feng, B. ER stress triggers MCP-1 expression through SET7/9-induced histone methylation in the kidneys of db/db mice. Am. J. Physiol. Ren. Physiol. 2014, 306, F916–F925. [Google Scholar] [CrossRef] [Green Version]
  31. Yuan, H.; Reddy, M.A.; Sun, G.; Lanting, L.; Wang, M.; Kato, M.; Natarajan, R. Involvement of p300/CBP and epigenetic histone acetylation in TGF-β1-mediated gene transcription in mesangial cells. Am. J. Physiol. Renal. Physiol. 2013, 304, F601–F613. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Siddiqi, F.S.; Majumder, S.; Thai, K.; Abdalla, M.; Hu, P.; Advani, S.L.; White, K.E.; Bowskill, B.B.; Guarna, G.; DosSantos, C.C.; et al. The Histone Methyltransferase Enzyme Enhancer of Zeste Homolog 2 Protects against Podocyte Oxidative Stress and Renal Injury in Diabetes. J. Am. Soc. Nephrol. 2016, 27, 2021–2034. [Google Scholar] [CrossRef] [Green Version]
  33. Liu, M.; Liang, K.; Zhen, J.; Zhou, M.; Wang, X.; Wang, Z.; Wei, X.; Zhang, Y.; Sun, Y.; Zhou, Z.; et al. Sirt6 deficiency exacerbates podocyte injury and proteinuria through targeting Notch signaling. Nat. Commun. 2017, 8, 413. [Google Scholar] [CrossRef] [Green Version]
  34. Lin, C.L.; Lee, P.H.; Hsu, Y.C.; Lei, C.C.; Ko, J.Y.; Chuang, P.C.; Huang, Y.T.; Wang, S.Y.; Wu, S.L.; Chen, Y.S.; et al. MicroRNA-29a promotion of nephrin acetylation ameliorates hyperglycemia-induced podocyte dysfunction. J. Am. Soc. Nephrol. 2014, 25, 1698–1709. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Kashihara, N.; Haruna, Y.; Kondeti, V.K.; Kanwar, Y.S. Oxidative stress in diabetic nephropathy. Curr. Med. Chem. 2010, 17, 4256–4269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Wei, P.Z.; Szeto, C.C. Mitochondrial dysfunction in diabetic kidney disease. Clin. Chim. Acta. 2019, 496, 108–116. [Google Scholar] [CrossRef] [PubMed]
  37. Qi, H.; Casalena, G.; Shi, S.; Yu, L.; Ebefors, K.; Sun, Y.; Zhang, W.; D’Agati, V.; Schlondorff, D.; Haraldsson, B.; et al. Glomerular Endothelial Mitochondrial Dysfunction Is Essential and Characteristic of Diabetic Kidney Disease Susceptibility. Diabetes 2017, 66, 763–778. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Ma, Y.; Chen, Z.; Tao, Y.; Zhu, J.; Yang, H.; Liang, W.; Ding, G. Increased mitochondrial fission of glomerular podocytes in diabetic nephropathy. Endocr. Connect. 2019, 8, 1206–1212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Hartleben, B.; Gödel, M.; Meyer-Schwesinger, C.; Liu, S.; Ulrich, T.; Köbler, S.; Wiech, T.; Grahammer, F.; Arnold, S.J.; Lindenmeyer, M.T.; et al. Autophagy influences glomerular disease susceptibility and maintains podocyte homeostasis in aging mice. J. Clin. Investig. 2010, 120, 1084–1096. [Google Scholar] [CrossRef] [Green Version]
  40. Hu, C.; Sun, L.; Xiao, L.; Han, Y.; Fu, X.; Xiong, X.; Xu, X.; Liu, Y.; Yang, S.; Liu, F.; et al. Insights into the Mechanisms Involved in the Expression and Regulation of Extracellular Matrix Proteins in Diabetic Nephropathy. Curr. Med. Chem. 2015, 22, 2858–2870. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Shirakawa, J.; Togashi, Y.; Tajima, K.; Orime, K.; Kikuchi, K.; Miyazaki, T.; Sato, K.; Kimura, M.; Goshima, Y.; Terauchi, Y. Plasminogen activator inhibitor-1 is associated with renal dysfunction independent of BMI and serum lipid levels in patients with type 2 diabetes. Diabetes Res. Clin. Pract. 2012, 97, e9–e12. [Google Scholar] [CrossRef]
  42. Satirapoj, B.; Dispan, R.; Radinahamed, P.; Kitiyakara, C. Urinary epidermal growth factor, monocyte chemoattractant protein-1 or their ratio as predictors for rapid loss of renal function in type 2 diabetic patients with diabetic kidney disease. BMC Nephrol. 2018, 19, 246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Wu, X.Q.; Zhang, D.D.; Wang, Y.N.; Tan, Y.Q.; Yu, X.Y.; Zhao, Y.Y. AGE/RAGE in diabetic kidney disease and ageing kidney. Free Radic. Biol. Med. 2021, 171, 260–271. [Google Scholar] [CrossRef] [PubMed]
  44. Kumar Pasupulati, A.; Chitra, P.S.; Reddy, G.B. Advanced glycation end products mediated cellular and molecular events in the pathology of diabetic nephropathy. Biomol. Concepts 2016, 7, 293–309. [Google Scholar] [CrossRef] [Green Version]
  45. Sharma, K.; Ziyadeh, F.N. Hyperglycemia and diabetic kidney disease. The case for transforming growth factor-beta as a key mediator. Diabetes 1995, 44, 1139–1146. [Google Scholar] [CrossRef] [PubMed]
  46. Ziyadeh, F.N. Mediators of diabetic renal disease: The case for tgf-Beta as the major mediator. J. Am. Soc. Nephrol. 2004, 15 (Suppl. S1), S55–S57. [Google Scholar] [CrossRef] [Green Version]
  47. Chen, S.; Jim, B.; Ziyadeh, F.N. Diabetic nephropathy and transforming growth factor-beta: Transforming our view of glomerulosclerosis and fibrosis build-up. Semin. Nephrol. 2003, 23, 532–543. [Google Scholar] [CrossRef] [PubMed]
  48. Chen, L.; Yang, T.; Lu, D.W.; Zhao, H.; Feng, Y.L.; Chen, H.; Chen, D.Q.; Vaziri, N.D.; Zhao, Y.Y. Central role of dysregulation of TGF-β/Smad in CKD progression and potential targets of its treatment. Biomed. Pharmacother. 2018, 101, 670–681. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Loboda, A.; Sobczak, M.; Jozkowicz, A.; Dulak, J. TGF-β1/Smads and miR-21 in Renal Fibrosis and Inflammation. Mediat. Inflamm. 2016, 2016, 8319283. [Google Scholar] [CrossRef] [Green Version]
  50. Schiffer, M.; Bitzer, M.; Roberts, I.S.; Kopp, J.B.; ten Dijke, P.; Mundel, P.; Böttinger, E.P. Apoptosis in podocytes induced by TGF-beta and Smad7. J. Clin. Investig. 2001, 108, 807–816. [Google Scholar] [CrossRef] [PubMed]
  51. Zhou, D.; Liu, Y. Renal fibrosis in 2015: Understanding the mechanisms of kidney fibrosis. Nat. Rev. Nephrol. 2016, 12, 68–70. [Google Scholar] [CrossRef] [PubMed]
  52. Loh, C.Y.; Chai, J.Y.; Tang, T.F.; Wong, W.F.; Sethi, G.; Shanmugam, M.K.; Chong, P.P.; Looi, C.Y. The E-Cadherin and N-Cadherin Switch in Epithelial-to-Mesenchymal Transition: Signaling, Therapeutic Implications, and Challenges. Cells 2019, 8, 1118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Li, J.; Qu, X.; Yao, J.; Caruana, G.; Ricardo, S.D.; Yamamoto, Y.; Yamamoto, H.; Bertram, J.F. Blockade of endothelial-mesenchymal transition by a Smad3 inhibitor delays the early development of streptozotocin-induced diabetic nephropathy. Diabetes 2010, 59, 2612–2624. [Google Scholar] [CrossRef] [Green Version]
  54. Peng, H.; Li, Y.; Wang, C.; Zhang, J.; Chen, Y.; Chen, W.; Cao, J.; Wang, Y.; Hu, Z.; Lou, T. ROCK1 Induces Endothelial-to-Mesenchymal Transition in Glomeruli to Aggravate Albuminuria in Diabetic Nephropathy. Sci. Rep. 2016, 6, 20304. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Kouzarides, T. Chromatin modifications and their function. Cell 2007, 128, 693–705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Tessarz, P.; Kouzarides, T. Histone core modifications regulating nucleosome structure and dynamics. Nat. Rev. Mol. Cell Biol. 2014, 15, 703–708. [Google Scholar] [CrossRef] [PubMed]
  57. Lawrence, M.; Daujat, S.; Schneider, R. Lateral Thinking: How Histone Modifications Regulate Gene Expression. Trends Genet. 2016, 32, 42–56. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Zhou, V.W.; Goren, A.; Bernstein, B.E. Charting histone modifications and the functional organization of mammalian genomes. Nat. Rev. Genet. 2011, 12, 7–18. [Google Scholar] [CrossRef] [PubMed]
  59. Miao, F.; Smith, D.D.; Zhang, L.; Min, A.; Feng, W.; Natarajan, R. Lymphocytes from patients with type 1 diabetes display a distinct profile of chromatin histone H3 lysine 9 dimethylation: An epigenetic study in diabetes. Diabetes 2008, 57, 3189–3198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Miao, F.; Gonzalo, I.G.; Lanting, L.; Natarajan, R. In vivo chromatin remodeling events leading to inflammatory gene transcription under diabetic conditions. J. Biol. Chem. 2004, 279, 18091–18097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Reddy, M.A.; Natarajan, R. Recent developments in epigenetics of acute and chronic kidney diseases. Kidney Int. 2015, 88, 250–261. [Google Scholar] [CrossRef] [PubMed]
  62. Zheng, W.; Guo, J.; Liu, Z.S. Effects of metabolic memory on inflammation and fibrosis associated with diabetic kidney disease: An epigenetic perspective. Clin. Epigenet. 2021, 13, 87. [Google Scholar] [CrossRef]
  63. Wang, X.; Liu, J.; Zhen, J.; Zhang, C.; Wan, Q.; Liu, G.; Wei, X.; Zhang, Y.; Wang, Z.; Han, H.; et al. Histone deacetylase 4 selectively contributes to podocyte injury in diabetic nephropathy. Kidney Int. 2014, 86, 712–725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Liang, T.; Qi, C.; Lai, Y.; Xie, J.; Wang, H.; Zhang, L.; Lin, T.; Jv, M.; Li, J.; Wang, Y.; et al. HDAC6-mediated α-tubulin deacetylation suppresses autophagy and enhances motility of podocytes in diabetic nephropathy. J. Cell. Mol. Med. 2020, 24, 11558–11572. [Google Scholar] [CrossRef]
  65. Liu, R.; Zhong, Y.; Li, X.; Chen, H.; Jim, B.; Zhou, M.M.; Chuang, P.Y.; He, J.C. Role of transcription factor acetylation in diabetic kidney disease. Diabetes 2014, 63, 2440–2453. [Google Scholar] [CrossRef] [Green Version]
  66. Li, X.; Li, C.; Sun, G. Histone Acetylation and Its Modifiers in the Pathogenesis of Diabetic Nephropathy. J. Diabetes Res. 2016, 2016, 4065382. [Google Scholar] [CrossRef] [Green Version]
  67. Yang, X.J.; Seto, E. HATs and HDACs: From structure, function and regulation to novel strategies for therapy and prevention. Oncogene 2007, 26, 5310–5318. [Google Scholar] [CrossRef]
  68. Sayyed, S.G.; Gaikwad, A.B.; Lichtnekert, J.; Kulkarni, O.; Eulberg, D.; Klussmann, S.; Tikoo, K.; Anders, H.J. Progressive glomerulosclerosis in type 2 diabetes is associated with renal histone H3K9 and H3K23 acetylation, H3K4 dimethylation and phosphorylation at serine 10. Nephrol. Dial. Transplant. 2010, 25, 1811–1817. [Google Scholar] [CrossRef] [PubMed]
  69. Huang, J.; Wan, D.; Li, J.; Chen, H.; Huang, K.; Zheng, L. Histone acetyltransferase PCAF regulates inflammatory molecules in the development of renal injury. Epigenetics 2015, 10, 62–72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Lazar, A.G.; Vlad, M.L.; Manea, A.; Simionescu, M.; Manea, S.A. Activated Histone Acetyltransferase p300/CBP-Related Signalling Pathways Mediate Up-Regulation of NADPH Oxidase, Inflammation, and Fibrosis in Diabetic Kidney. Antioxidants 2021, 10, 1356. [Google Scholar] [CrossRef]
  71. Wang, Y.; Wang, Y.; Luo, M.; Wu, H.; Kong, L.; Xin, Y.; Cui, W.; Zhao, Y.; Wang, J.; Liang, G.; et al. Novel curcumin analog C66 prevents diabetic nephropathy via JNK pathway with the involvement of p300/CBP-mediated histone acetylation. Biochim. Biophys. Acta 2015, 1852, 34–46. [Google Scholar] [CrossRef] [Green Version]
  72. Deb, D.K.; Chen, Y.; Sun, J.; Wang, Y.; Li, Y.C. ATP-citrate lyase is essential for high glucose-induced histone hyperacetylation and fibrogenic gene upregulation in mesangial cells. Am. J. Physiol. Renal. Physiol. 2017, 313, F423–F429. [Google Scholar] [CrossRef] [Green Version]
  73. Muraoka, H.; Hasegawa, K.; Sakamaki, Y.; Minakuchi, H.; Kawaguchi, T.; Yasuda, I.; Kanda, T.; Tokuyama, H.; Wakino, S.; Itoh, H. Role of Nampt-Sirt6 Axis in Renal Proximal Tubules in Extracellular Matrix Deposition in Diabetic Nephropathy. Cell Rep. 2019, 27, 199–212.e5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Chen, S.; Feng, B.; George, B.; Chakrabarti, R.; Chen, M.; Chakrabarti, S. Transcriptional coactivator p300 regulates glucose-induced gene expression in endothelial cells. Am. J. Physiol. Endocrinol. Metab. 2010, 298, E127–E137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Gaikwad, A.B.; Gupta, J.; Tikoo, K. Epigenetic changes and alteration of Fbn1 and Col3A1 gene expression under hyperglycaemic and hyperinsulinaemic conditions. Biochem. J. 2010, 432, 333–341. [Google Scholar] [CrossRef] [Green Version]
  76. Xu, H.; Wu, X.; Qin, H.; Tian, W.; Chen, J.; Sun, L.; Fang, M.; Xu, Y. Myocardin-Related Transcription Factor A Epigenetically Regulates Renal Fibrosis in Diabetic Nephropathy. J. Am. Soc. Nephrol. 2015, 26, 1648–1660. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Hasegawa, K.; Wakino, S.; Simic, P.; Sakamaki, Y.; Minakuchi, H.; Fujimura, K.; Hosoya, K.; Komatsu, M.; Kaneko, Y.; Kanda, T.; et al. Renal tubular Sirt1 attenuates diabetic albuminuria by epigenetically suppressing Claudin-1 overexpression in podocytes. Nat. Med. 2013, 19, 1496–1504. [Google Scholar] [CrossRef] [Green Version]
  78. Yuan, H.; Reddy, M.A.; Deshpande, S.; Jia, Y.; Park, J.T.; Lanting, L.L.; Jin, W.; Kato, M.; Xu, Z.G.; Das, S.; et al. Epigenetic Histone Modifications Involved in Profibrotic Gene Regulation by 12/15-Lipoxygenase and Its Oxidized Lipid Products in Diabetic Nephropathy. Antioxid. Redox Signal. 2016, 24, 361–375. [Google Scholar] [CrossRef] [PubMed]
  79. Noh, H.; Oh, E.Y.; Seo, J.Y.; Yu, M.R.; Kim, Y.O.; Ha, H.; Lee, H.B. Histone deacetylase-2 is a key regulator of diabetes- and transforming growth factor-beta1-induced renal injury. Am. J. Physiol. Renal. Physiol. 2009, 297, F729–F739. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Yoshikawa, M.; Hishikawa, K.; Marumo, T.; Fujita, T. Inhibition of histone deacetylase activity suppresses epithelial-to-mesenchymal transition induced by TGF-beta1 in human renal epithelial cells. J. Am. Soc. Nephrol. 2007, 18, 58–65. [Google Scholar] [CrossRef] [Green Version]
  81. Sun, X.Y.; Qin, H.J.; Zhang, Z.; Xu, Y.; Yang, X.C.; Zhao, D.M.; Li, X.N.; Sun, L.K. Valproate attenuates diabetic nephropathy through inhibition of endoplasmic reticulum stress-induced apoptosis. Mol. Med. Rep. 2016, 13, 661–668. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. De Marinis, Y.; Cai, M.; Bompada, P.; Atac, D.; Kotova, O.; Johansson, M.E.; Garcia-Vaz, E.; Gomez, M.F.; Laakso, M.; Groop, L. Epigenetic regulation of the thioredoxin-interacting protein (TXNIP) gene by hyperglycemia in kidney. Kidney Int. 2016, 89, 342–353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Malek, V.; Sharma, N.; Gaikwad, A.B. Histone Acetylation Regulates Natriuretic Peptides and Neprilysin Gene Expressions in Diabetic Cardiomyopathy and Nephropathy. Curr. Mol. Pharmacol. 2019, 12, 61–71. [Google Scholar] [CrossRef]
  84. Ma, X.; Wang, Q. Short-Chain Fatty Acids Attenuate Renal Fibrosis and Enhance Autophagy of Renal Tubular Cells in Diabetic Mice Through the HDAC2/ULK1 Axis. Endocrinol. Metab. 2022, 37, 432–443. [Google Scholar] [CrossRef] [PubMed]
  85. Greer, E.L.; Shi, Y. Histone methylation: A dynamic mark in health, disease and inheritance. Nat. Rev. Genet. 2012, 13, 343–357. [Google Scholar] [CrossRef] [Green Version]
  86. Kato, M.; Natarajan, R. Epigenetics and epigenomics in diabetic kidney disease and metabolic memory. Nat. Rev. Nephrol. 2019, 15, 327–345. [Google Scholar] [CrossRef] [PubMed]
  87. Black, J.C.; VanRechem, C.; Whetstine, J.R. Histone lysine methylation dynamics: Establishment, regulation, and biological impact. Mol. Cell 2012, 48, 491–507. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Hyun, K.; Jeon, J.; Park, K.; Kim, J. Writing, erasing and reading histone lysine methylations. Exp. Mol. Med. 2017, 49, e324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Klose, R.J.; Zhang, Y. Regulation of histone methylation by demethylimination and demethylation. Nat. Rev. Mol. Cell Biol. 2007, 8, 307–318. [Google Scholar] [CrossRef]
  90. Chen, H.; Huang, Y.; Zhu, X.; Liu, C.; Yuan, Y.; Su, H.; Zhang, C.; Liu, C.; Xiong, M.; Qu, Y.; et al. Histone demethylase UTX is a therapeutic target for diabetic kidney disease. J. Physiol. 2019, 597, 1643–1660. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Lin, C.L.; Hsu, Y.C.; Huang, Y.T.; Shih, Y.H.; Wang, C.J.; Chiang, W.C.; Chang, P.J. A KDM6A-KLF10 reinforcing feedback mechanism aggravates diabetic podocyte dysfunction. EMBO Mol. Med. 2019, 11, e9828. [Google Scholar] [CrossRef]
  92. Bai, S.; Xiong, X.; Tang, B.; Ji, T.; Li, X.; Qu, X.; Li, W. hsa-miR-199b-3p Prevents the Epithelial-Mesenchymal Transition and Dysfunction of the Renal Tubule by Regulating E-cadherin through Targeting KDM6A in Diabetic Nephropathy. Oxidative Med. Cell. Longev. 2021, 2021, 8814163. [Google Scholar] [CrossRef]
  93. Majumder, S.; Thieme, K.; Batchu, S.N.; Alghamdi, T.A.; Bowskill, B.B.; Kabir, M.G.; Liu, Y.; Advani, S.L.; White, K.E.; Geldenhuys, L.; et al. Shifts in podocyte histone H3K27me3 regulate mouse and human glomerular disease. J. Clin. Investig. 2018, 128, 483–499. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Cai, M.; Bompada, P.; Atac, D.; Laakso, M.; Groop, L.; De Marinis, Y. Epigenetic regulation of glucose-stimulated osteopontin (OPN) expression in diabetic kidney. Biochem. Biophys. Res. Commun. 2016, 469, 108–113. [Google Scholar] [CrossRef] [PubMed]
  95. Zhang, L.; Chen, L.; Gao, C.; Chen, E.; Lightle, A.R.; Foulke, L.; Zhao, B.; Higgins, P.J.; Zhang, W. Loss of Histone H3 K79 Methyltransferase Dot1l Facilitates Kidney Fibrosis by Upregulating Endothelin 1 through Histone Deacetylase 2. J. Am. Soc. Nephrol. 2020, 31, 337–349. [Google Scholar] [CrossRef] [PubMed]
  96. Das, F.; Bera, A.; Ghosh-Choudhury, N.; Sataranatarajan, K.; Kamat, A.; Kasinath, B.S.; Choudhury, G.G. High glucose-stimulated enhancer of zeste homolog-2 (EZH2) forces suppression of deptor to cause glomerular mesangial cell pathology. Cell. Signal. 2021, 86, 110072. [Google Scholar] [CrossRef]
  97. Fang, N.; Li, P. O-linked N-acetylglucosaminyltransferase OGT inhibits diabetic nephropathy by stabilizing histone methyltransferases EZH2 via the HES1/PTEN axis. Life Sci. 2021, 274, 119226. [Google Scholar] [CrossRef] [PubMed]
  98. Aufhauser, D.D., Jr.; Hernandez, P.; Concors, S.J.; O’Brien, C.; Wang, Z.; Murken, D.R.; Samanta, A.; Beier, U.H.; Krumeich, L.; Bhatti, T.R.; et al. HDAC2 targeting stabilizes the CoREST complex in renal tubular cells and protects against renal ischemia/reperfusion injury. Sci. Rep. 2021, 11, 9018. [Google Scholar] [CrossRef] [PubMed]
  99. Wang, J.; Yan, W.; Peng, X.; Jiang, Y.; He, L.; Peng, Y.; Chen, X.; Ye, M.; Zhuo, H. Functional Role of SUV39H1 in Human Renal Tubular Epithelial Cells under High-glucose Ambiance. Inflammation 2018, 41, 1–10. [Google Scholar] [CrossRef] [PubMed]
  100. Wang, J.; Shen, X.; Liu, J.; Chen, W.; Wu, F.; Wu, W.; Meng, Z.; Zhu, M.; Miao, C. High glucose mediates NLRP3 inflammasome activation via upregulation of ELF3 expression. Cell Death Dis. 2020, 11, 383. [Google Scholar] [CrossRef]
  101. Qi, J.; Wu, Q.; Cheng, Q.; Chen, X.; Zhu, M.; Miao, C. High Glucose Induces Endothelial COX2 and iNOS Expression via Inhibition of Monomethyltransferase SETD8 Expression. J. Diabetes Res. 2020, 2020, 2308520. [Google Scholar] [CrossRef] [PubMed]
  102. Paneni, F.; Costantino, S.; Battista, R.; Castello, L.; Capretti, G.; Chiandotto, S.; Scavone, G.; Villano, A.; Pitocco, D.; Lanza, G.; et al. Adverse epigenetic signatures by histone methyltransferase Set7 contribute to vascular dysfunction in patients with type 2 diabetes mellitus. Circ. Cardiovasc. Genet. 2015, 8, 150–158. [Google Scholar] [CrossRef] [PubMed]
  103. Jia, Y.; Reddy, M.A.; Das, S.; Oh, H.J.; Abdollahi, M.; Yuan, H.; Zhang, E.; Lanting, L.; Wang, M.; Natarajan, R. Dysregulation of histone H3 lysine 27 trimethylation in transforming growth factor-β1-induced gene expression in mesangial cells and diabetic kidney. J. Biol. Chem. 2019, 294, 12695–12707. [Google Scholar] [CrossRef] [PubMed]
  104. Lu, L.; Li, X.; Zhong, Z.; Zhou, W.; Zhou, D.; Zhu, M.; Miao, C. KMT5A downregulation participated in High Glucose-mediated EndMT via Upregulation of ENO1 Expression in Diabetic Nephropathy. Int. J. Biol. Sci. 2021, 17, 4093–4107. [Google Scholar] [CrossRef]
  105. Lin, S.H.; Ho, W.T.; Wang, Y.T.; Chuang, C.T.; Chuang, L.Y.; Guh, J.Y. Histone methyltransferase Suv39h1 attenuates high glucose-induced fibronectin and p21(WAF1) in mesangial cells. Int. J. Biochem. Cell Biol. 2016, 78, 96–105. [Google Scholar] [CrossRef]
  106. Komers, R.; Mar, D.; Denisenko, O.; Xu, B.; Oyama, T.T.; Bomsztyk, K. Epigenetic changes in renal genes dysregulated in mouse and rat models of type 1 diabetes. Lab. Investig. 2013, 93, 543–552. [Google Scholar] [CrossRef] [Green Version]
  107. Mimura, I. Epigenetic memory in kidney diseases. Kidney Int. 2016, 89, 274–277. [Google Scholar] [CrossRef]
  108. Zeng, S.; Wu, X.; Chen, X.; Xu, H.; Zhang, T.; Xu, Y. Hypermethylated in cancer 1 (HIC1) mediates high glucose induced ROS accumulation in renal tubular epithelial cells by epigenetically repressing SIRT1 transcription. Biochim. Biophys. Acta Gene Regul. Mech. 2018, 1861, 917–927. [Google Scholar] [CrossRef] [PubMed]
  109. Guo, Q.; Li, X.; Han, H.; Li, C.; Liu, S.; Gao, W.; Sun, G. Histone Lysine Methylation in TGF-β1 Mediated p21 Gene Expression in Rat Mesangial Cells. Biomed. Res. Int. 2016, 2016, 6927234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Li, X.; Li, C.; Li, X.; Cui, P.; Li, Q.; Guo, Q.; Han, H.; Liu, S.; Sun, G. Involvement of Histone Lysine Methylation in p21 Gene Expression in Rat Kidney In Vivo and Rat Mesangial Cells In Vitro under Diabetic Conditions. J. Diabetes Res. 2016, 2016, 3853242. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Fan, Y.; Yang, Q.; Yang, Y.; Gao, Z.; Ma, Y.; Zhang, L.; Liang, W.; Ding, G. Sirt6 Suppresses High Glucose-Induced Mitochondrial Dysfunction and Apoptosis in Podocytes through AMPK Activation. Int. J. Biol. Sci. 2019, 15, 701–713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Fan, Y.; Fan, H.; Li, P.; Liu, Q.; Huang, L.; Zhou, Y. Mitogen-activating protein kinase-3, inhibited by Astragaloside IV through H3 lysine 4 monomethylation, promotes the progression of diabetic nephropathy by inducing apoptosis. Bioengineered 2022, 13, 11517–11529. [Google Scholar] [CrossRef]
  113. Goru, S.K.; Kadakol, A.; Pandey, A.; Malek, V.; Sharma, N.; Gaikwad, A.B. Histone H2AK119 and H2BK120 mono-ubiquitination modulate SET7/9 and SUV39H1 in type 1 diabetes-induced renal fibrosis. Biochem. J. 2016, 473, 3937–3949. [Google Scholar] [CrossRef] [PubMed]
  114. Kuo, F.C.; Chao, C.T.; Lin, S.H. The Dynamics and Plasticity of Epigenetics in Diabetic Kidney Disease: Therapeutic Applications Vis-à-Vis. Int. J. Mol. Sci. 2022, 23, 843. [Google Scholar] [CrossRef] [PubMed]
  115. Fox, C.S.; Matsushita, K.; Woodward, M.; Bilo, H.J.; Chalmers, J.; Heerspink, H.J.; Lee, B.J.; Perkins, R.M.; Rossing, P.; Sairenchi, T.; et al. Associations of kidney disease measures with mortality and end-stage renal disease in individuals with and without diabetes: A meta-analysis. Lancet 2012, 380, 1662–1673. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Cobo, I.; Tanaka, T.N.; Chandra Mangalhara, K.; Lana, A.; Yeang, C.; Han, C.; Schlachetzki, J.; Challcombe, J.; Fixsen, B.R.; Sakai, M.; et al. DNA methyltransferase 3 alpha and TET methylcytosine dioxygenase 2 restrain mitochondrial DNA-mediated interferon signaling in macrophages. Immunity 2022, 55, 1386–1401.e10. [Google Scholar] [CrossRef]
  117. Dawoud, A.A.Z.; Gilbert, R.D.; Tapper, W.J.; Cross, N.C.P. Clonal myelopoiesis promotes adverse outcomes in chronic kidney disease. Leukemia 2022, 36, 507–515. [Google Scholar] [CrossRef] [PubMed]
Table 1. Major risk factors for kidney disease in patients with diabetes mellitus and their management.
Table 1. Major risk factors for kidney disease in patients with diabetes mellitus and their management.
Risk FactorTargetRecommended Management
HyperglycemiaHbA1c < 7% (<6.5% if achievable without a substantially increased risk for hypoglycemia, or 7.5–8.0% in patients with substantial comorbidities or reduced life expectancy)Sodium glucose transporter inhibitors
Other antidiabetic agents if needed
HypertensionSystolic blood pressure < 140 mmHg (<120 mmHg if it can be reasonably achieved)
Diastolic blood pressure 80–90 mmHg
Angiotensin converting enzyme inhibitors or angiotensin receptor blockers
Finerenone
Other antihypertensive agents if needed
ObesityAt least 5% weight loss from baselinePhysical activity and low-calorie diet (500 kcal/day deficit)
DyslipidemiaLow density lipoprotein cholesterol levels < 55 mg/dl and LDL-C reduction > 50% from baselineStatins
Ezetimibe and proprotein convertase subtilisin/kexin type 9 inhibitors if needed
Table 2. Principal mechanisms implicated in the pathogenesis of diabetic kidney disease.
Table 2. Principal mechanisms implicated in the pathogenesis of diabetic kidney disease.
MechanismAssociated MechanismsRegulators
Impaired renal hemodynamicsHyperglycemia, hyperaminoacidemia, hypertension and obesity [13]ET-1 [14]
Oxidative stressInflammation and renal fibrosis [15]Txnip [16,17]
Dysregulation of the immune system/
Inflammation
Immune cells and adhesion molecules [18,19]NF-κB [20], NLRP3 [21]
Compromised autophagyAGEs, oxidative stress, ER stress, RAAS activation and renal fibrosis [22,23]mTORC1 [22], insulin [24]
FibrosisOxidative stress [15], autophagy [22,23]TGF-β, MAPK, Wnt/β-catenin, PI3K/Akt, JAK/STAT, and Notch pathways [25], PAI-1 [26]
Histone modificationFibrosis [27,28], inflammation [29,30], hypertrophy [31,32], apoptosis [33,34]
ET-1: endothelin 1, Txnip: thioredoxin interacting protein, NF-κB: nuclear factor kappa-light-chain-enhancer of activated B cells, NLRP3: nucleotide binding and oligomerization domain-like receptor family pyrin domain-containing 3, AGEs: advanced glycation end-products, ER: endoplasmic reticulum, RAAS: renin–angiotensin–aldosterone system, TGF-β: transforming growth factor β, MAPK: mitogen-activated protein kinase, PI3K: phosphatidyl-inositol-3 kinase, PAI-1: plasminogen activator inhibitor-1.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kourtidou, C.; Tziomalos, K. The Role of Histone Modifications in the Pathogenesis of Diabetic Kidney Disease. Int. J. Mol. Sci. 2023, 24, 6007. https://doi.org/10.3390/ijms24066007

AMA Style

Kourtidou C, Tziomalos K. The Role of Histone Modifications in the Pathogenesis of Diabetic Kidney Disease. International Journal of Molecular Sciences. 2023; 24(6):6007. https://doi.org/10.3390/ijms24066007

Chicago/Turabian Style

Kourtidou, Christodoula, and Konstantinos Tziomalos. 2023. "The Role of Histone Modifications in the Pathogenesis of Diabetic Kidney Disease" International Journal of Molecular Sciences 24, no. 6: 6007. https://doi.org/10.3390/ijms24066007

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