Next Article in Journal
Zinc Oxide Nanoparticles Protected with Terpenoids as a Substance in Redox Imbalance Normalization in Burns
Next Article in Special Issue
Neuroprotective Effects of Tripeptides—Epigenetic Regulators in Mouse Model of Alzheimer’s Disease
Previous Article in Journal
The Antiviral Role of Galectins toward Influenza A Virus Infection—An Alternative Strategy for Influenza Therapy
Previous Article in Special Issue
m6A RNA Methylation in Systemic Autoimmune Diseases—A New Target for Epigenetic-Based Therapy?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Epigenetic Mechanisms Involved in Cisplatin-Induced Nephrotoxicity: An Update

1
Center of Molecular Biology and Pharmacogenetics, Scientific and Technological Bioresource Nucleus, Universidad de La Frontera, Temuco 4811230, Chile
2
Department of Medical Technology, Faculty of Medicine, Universidad de Chile, Santiago 8380453, Chile
3
Faculty of Pharmaceutical Sciences, University of Campinas, Campinas 13083970, SP, Brazil
*
Author to whom correspondence should be addressed.
Pharmaceuticals 2021, 14(6), 491; https://0-doi-org.brum.beds.ac.uk/10.3390/ph14060491
Submission received: 1 May 2021 / Revised: 19 May 2021 / Accepted: 20 May 2021 / Published: 21 May 2021
(This article belongs to the Special Issue Epigenetic Drugs)

Abstract

:
Cisplatin is an antineoplastic drug used for the treatment of many solid tumors. Among its various side effects, nephrotoxicity is the most detrimental. In recent years, epigenetic regulation has emerged as a modulatory mechanism of cisplatin-induced nephrotoxicity, involving non-coding RNAs, DNA methylation and histone modifications. These epigenetic marks alter different signaling pathways leading to damage and cell death. In this review, we describe how different epigenetic modifications alter different pathways leading to cell death by apoptosis, autophagy, necroptosis, among others. The study of epigenetic regulation is still under development, and much research remains to fully determine the epigenetic mechanisms underlying cell death, which will allow leading new strategies for the diagnosis and therapy of this disease.

1. Introduction

Globally, cancer is the second leading cause of death. In 2018, 18.1 million people worldwide had cancer, and 9.6 million died from this disease [1]. Cisplatin [cis-diamminedichloroplatinum (II), CDDP] is a well-known chemotherapeutic drug used for the treatment of numerous human cancer in solid organs, including head and neck [2], testis [3], small cells [4] and non-small cells [5] lung cancer, ovarian [6], cervical [7] and bladder [8]. Regarding the treatment of cancer cells, CDDP is potentially accompanied by some side effects such as ototoxicity, gastrotoxicity, neurotoxicity, myelosuppression, and allergic reactions [9]. Still, the main limiting side effect of CDDP use is nephrotoxicity [10]. Nephrotoxicity is defined as the rapid deterioration in kidney function due to the toxic effect of drugs, medications, and/or chemicals [11]. Cisplatin-induced-nephrotoxicity can occur in several ways, but the most severe and common is acute kidney injury, whose incidence after a single dose of CDDP fluctuates between 20–30% in patients who receive an initial dose of 50–100 mg/m2 [9]. Cisplatin is concentrated and reabsorbed by tubular renal cells (5 times more than in the blood), which triggers a rapid decrease in renal function [12], resulting in morphological changes and altered signaling pathways, such as vacuolation, mitochondrial dysfunction, cell cycle arrest and apoptosis [13,14]. The diagnosis of kidney damage is carried out using standard biomarkers, such as serum creatinine (sCr) and blood urea nitrogen (BUN). However, their sensitivity and specificity are very low, and values are altered when significant kidney damage already exists [15].
In recent years, several studies showed that epigenetic processes are involved in cisplatin-induced nephrotoxicity. Wu and Morris [16] defined epigenetic as the “study of the changes in gene expression, which occur in organisms with differentiated cells, and the mitotic inheritance of given patterns of gene expression”. These modifications regulate gene expression without changing the primary DNA sequence, triggering activation of transcription or gene silencing [17]. The main epigenetic mechanisms are represented by DNA methylation, histone modifications, and non-coding RNA. In general, epigenetic modifications are stable and heritable during cell divisions; however, they are potentially reversible and can be influenced by environmental factors. This review discusses recent findings in the epigenetics landscape of cisplatin-induced nephrotoxicity, focusing on the three mechanisms mentioned above.

2. Induction Mechanism Cisplatin-Induced Nephrotoxicity

Despite its therapeutic efficacy, the use of CDDP is limited due to the manifestation of side effects in normal tissues, mainly nephrotoxicity, whose incidence fluctuates between 20–30% in patients undergoing CDDP chemotherapy [9].
Cisplatin exerts its cytotoxic consequences by reacting with DNA, which ultimately culminates in irreversible apoptosis. Once CDDP goes through the cytosol, the low concentration of chloride present triggers the hydrolysis of the Pt-Cl bonds. One of the Cl ligands is replaced by H2O to form the first positively charged CDDP derivative [18]. The subsequent hydrolysis results in the replacement of the next Cl ligand, leading to the formation of the second CDDP derivative [18]. Cisplatin hydrolyzed complexes bind to negatively charged DNA bases, inducing DNA damage by forming different types of adducts, resulting in the arrest of the cell cycle. Cisplatin interacts primarily with the N7 sites of purine residues in DNA to form inter-strand DNA crosslinks and intra-strand crosslinks at 5′-GG-3′ sites of DNA bases. This results in defective DNA strands, causing the arrest of DNA synthesis and replication [19]. The main molecular sensor of DNA damage produced by CDPP is ATR (ataxia telangiectasia and Rad3-related), which is recruited to the site of damage and is co-localized with H2AX, forming nuclear foci. This is continued by the recruitment and activation of protein kinases, such as Chk1 and Chk2, which phosphorylate p53 [20], activating the signaling cascade that leads to cell death by apoptosis [21,22,23,24,25,26] or necrosis [27,28,29]. Although this way of DNA damage occurs in all cells, especially those that divide rapidly, such as tumor cells, this is considered the main mechanism mediating CDDP antitumor activity. CDDP cytotoxicity is not limited to cancer cells. In normal cells may lead to dangerous adverse effects such as ototoxicity, gastrotoxicity, myelosuppression, and allergic reactions. However, the major negative side effect of CDDP corresponds to nephrotoxicity [10]. After a single dose of CDDP (50–100 mg/m2), approximately 30% of patients develop this pathology [9]. The nephrotoxic effect is produced by CDDP accumulation in the kidney, mainly in the S3 segment of the proximal tubules [30]. One of the reasons for its accumulation may be determined by the high density of negatively charged mitochondria in the proximal tubular cells, which attracts positively charged CDDP hydrolyzed complexes [31]. Once CDDP enters the cell, it can induce various mechanisms resulting in cell damage. The intrinsic pathway is the main apoptotic pathway during CDDP-induced nephrotoxicity. Cisplatin decreases anti-apoptotic proteins such as Bcl-2, Bcl-XL, and Mcl-1 [32,33]. This triggers external membrane permeabilization of the mitochondria, releasing apoptotic factors, such as cytochrome c [34]. Once cytochrome c is released, recruitment and activation of caspase-9 occurs, therefore activating caspase-3, initiating the process of caspase-dependent apoptosis [35]. Cisplatin treatment also activates p53 [36,37,38,39], which phosphorylation results in the activation of several molecules, such as p53-up-regulated modulator of apoptosis (PUMA-α) and P53-induced protein with a death domain (PIDD) [21]. PIDD activates caspase-2, releasing apoptosis-inducing factor (AIF) from the mitochondria, resulting in caspase-independent activation [19]. On the other hand, PUMA-α translocates into mitochondria and interacts with BCL-XL, while BAX and BAK finally produce the release of cytochrome c, leading to caspase-dependent apoptosis [32,33]. Additionally, caspase-6 and -7 induction are produced during CDDP treatment [40]. Cisplatin also activates three major MAPKs: extracellular signal-regulated kinase (ERK), p38, and c-JUN N-terminal kinase (JNK) [35,41]. This induces oxidative stress, activating TNF-α and the consequent inflammatory response during CDDP-induced nephrotoxicity [42]. The augmented expression of ERK1/2 by CDDP treatment also activates NF-κβ and p53 signaling [43]. TNF-α increases tumor necrosis factors 1 and 2, TNFR1 and TNFR2, triggering the extrinsic apoptosis pathway through the activation of caspase-8, which will eventually lead to caspase-3 activation [9]. Cisplatin also regulates p21; its up-regulation protects from nephrotoxic damage [21]. Oxidative stress is a key factor contributing to nephrotoxicity since CDDP can react with glutathione, decreasing non-enzymatic (GSH and NADPH) and enzymatic (SOD, CAT, GPx, among others) antioxidant defense mechanisms [44]. Cisplatin also affects the mitochondrial respiratory complexes, leading to the inhibition of complexes I to IV of the respiratory chain and decreasing intracellular ATP levels [45]. In addition, CDDP also reduces the number of mitochondria in normal renal cells [43]. Finally, CDDP causes the activation of an adaptative program in the endoplasmic reticulum (ER) known as unfolded protein response (UPR), which inositol-requiring enzyme-1 (IRE1), double-stranded RNA-activated protein kinase-like ER kinase (PERK), and activating transcription factor-6 (ATF6) are dissociated from ER chaperone GRP78, activating signal transduction to inhibit protein translation [46]. Therefore, cisplatin-induced nephrotoxicity is a multifactorial and complex process involving multiple signaling pathways that in turn, may also be regulated by epigenetic mechanisms.

3. DNA Methylation Role in Cisplatin-Induced Renal Dysfunction

DNA methylation is an epigenetic modification involving the covalent addition of a methyl group to carbon 5 of a cytosine at CpG dinucleotides through the action of DNA methyltransferase enzymes (DNMTs) [47], which include DNMT1, DNMT2, DNMT3A, DNMT3B, and DNMT3L, contributing to hypermethylated (high methylation levels) or hypomethylated (low methylation levels) states of genes or genomic regions containing CpG residues [48]. In mammals, 70–80% of CpG dinucleotides are methylated, whereas CpG islands in promoter regions of genes are demethylated, allowing transcription to proceed [49]. Therefore, the main function of DNA methylation is the regulation of transcription. Thus, hypomethylation in the promoter region correlates with gene activation by increased access for transcription factors, whereas hypermethylation in the promoter region results in loss of gene expression. Studies associating different DNA methylation patterns with CDDP-induced nephrotoxicity are scarce. Reports show that treatment of tubular cells with 5-aza-2′-deoxycytidine (5-AZA), a DNA methylation inhibitor, increases CDDP-induced apoptosis [50,51]. In addition, IRF8, a pro-apoptotic factor, is hypomethylated and induced after CDDP treatment, contributing to renal tubular cell apoptosis (Figure 1) [50]. In conjunction with 5-AZA, cisplatin decreases the expression of DNMT1, a maintenance DNMT, attenuating CDDP-induced nephrotoxicity [52]. It is recognized that OCT2, a transporter involved in organic cation transport, can transport CDDP to proximal tubules [53]. Based on this, studies show that methylation of the promoter region of OCT2 dramatically reduces the transcriptional activity of this transporter [54], which may contribute to increased toxicity.

4. Cisplatin Nephrotoxicity and Histone Modifications

The dynamic changes in chromatin structure, which allow its decondensation and remodeling, are processes necessary for gene transcription, DNA repair, and replication, carried out by post-translational modifications in the different histones [55]. Histones are the basic structural proteins of nucleosomes, which bind to double-helical DNA, forming the DNA-histone complex [47]. There are several types of histones, such as linker histones (H1 and H5) and core histones (H2A, H2B, H3, and H4), whose function is associated with DNA packaging [49], where CDDP can easily react with H1 at methionine and glutamate residues, forming tertiary complexes that prevent DNA repair and increase sensitivity to CDDP [56]. Histone modifications occur predominantly at the N-terminal end, changing chromatin structure, positively or negatively affecting gene expression [49]. There are at least 8 histone modifications, but the most studied include histone acetylation and methylation, mostly in lysine and arginine residues.
Histone methylation involves the addition of a methyl group to a basic amino acid (lysine, arginine, and rarely in histidine) of core histones by a group of enzymes called histone methyltransferases (HMTs) [49]. Generally, methylation at H3 lysine 4 (H3K4), H3K36, and H3K79 activate gene transcription, whereas methylation at H3K9, H3K27, and H4K20 is associated with transcriptional repression [47]. CDDP treatment increases H3K27 trimethylation, leading to increased caspase 3 and decreased cell viability (Figure 2) [35]. Additionally, inhibition of enhancer of zeste homolog 2 (EZH2) -a histone methyltransferase- with 3-deazaneplanocin A (DZNep) in rat proximal tubular cells aggravates apoptosis by decreasing mTOR complexes [57].
On the other hand, histone acetylation involves adding an acetyl group to the lysine residue of core histones by a group of enzymes called histone acetyltransferases (HATs) [47]. This promotes open chromatin because a negative charge is added to the positively charged lysines, reducing the strong DNA-histone interaction, and thus, gene expression is activated [58]. One member of HATs is p300. This HAT can directly bind to transcription factors, such as p53 or NF-κB, and regulate their activities by acetylation [59,60]. Cisplatin induces p300 activation, acetylating lysines 18, 27, and 9 on H3, triggering functional and histological injury, increasing oxidative stress, inflammation, and apoptosis (Figure 2) [61]. One strategy to decrease CDDP-induced damage has been the inhibition of p300 by garcinol, attenuating the oxidative stress, inflammation, and apoptosis, by reducing the acetylation of the p65 subunit of NF-κB [61].
Acetylation is a reversible process catalyzed by enzymes called histone deacetylases (HDACs) [49]. HDACs are subdivided into four types: class I (HDAC1, 2, 3, and 8), class IIa (HDAC4, 5, 7, and 9), class IIb (HDAC6 and 10), class III (SIRT1-7), and class IV (HDAC11) [58]. One of the strategies to decrease CDDP-induced nephrotoxicity has been the use of HDACs class II inhibitors, as their overexpression is associated with increased apoptosis [62,63]. Studies have reported that the level of acetylated histone H3 is modified after CDDP treatment, but CDDP does not alter the abundance of histone H4 [64]. However, controversial results have been reported concerning acetylation of histone H3 with CDDP treatment. On the one hand, CDDP reduces levels of histone H3 acetylation in the injured kidney tissue [46,65]; however, another study demonstrated an increase in histone H3 acetylation after CDDP treatment in human renal cortical epithelial (HRCE) cells [66]. Although HDACs inhibitors do not prevent the DNA adducts formation and the initial damage response, they can interfere with the subsequent signaling [67]. Inhibition of HDAC6 by N-hydroxy-4-(2-methoxy-5-(methyl(2-methylquinazolin-4-yl)amino)phenoxy)butanamide (23BB) reduces apoptosis and ER stress in the kidney of cisplatin-injured mice [46]. Additionally, HDAC6 inhibition by trichostatin A (TSA) reduces the renal pathological damage caused by CDDP treatment [65]. Finally, reduction of HDAC4 and HDAC5 with LMK-235 blocks the caspase-3 cleavage in HRCE cells [66]. The use of TSA, valproate (VPA) or suberoylanilide hydroxamic acid (SAHA) can confer kidney protection during CDDP treatment by H3K27 acetylation and IL-9 down-regulation [62], increase of BMP-7 [63], AMWAP [68,69] and PSTPIP2 [70]; increased CREB phosphorylation [71], reduction of caspase-3 activation [71] or by autophagy stimulation [72]. Even more, it has been demonstrated that SAHA inhibits and TSA delays phosphorylation, acetylation, and activation of p53 in rat proximal tubular cells [67]. On the other hand, overexpression of HDAC class III decreases CDDP-induced damage. Pharmacological overexpression of SIRT1 ameliorates the increased acetylation of p65 of NF-κB [73] and p53 [74] during cisplatin treatment, which decreases apoptosis, oxidative stress, and inflammation [64,75]. However, SIRT2 up-regulation increases phosphorylation of p38 and JNK, inducing apoptosis and renal damage in renal tissue [76]. In the same way, SIRT6 overexpression attenuates cisplatin-induced damage by repressing the expression of ERK1/2, thus inhibiting NF-κB and p53 signaling [43]. Finally, SIRT3 activation protects kidney tubular cells against oxidative stress by PARP1 inhibition [77], and its deficiency augments CDDP nephrotoxicity by increasing apoptosis and inflammatory response (Figure 2) [78]. However, silencing SIRT5 expression significantly improves kidney function and decreases tissue damage in mice submitted to CDDP damage [79].

5. Involvement of Non-Coding RNAs in Cisplatin Nephrotoxicity

Less than 2% of the human genome is transcribed into RNA to be translated into proteins. The remaining 98% is transcribed into a class of RNA that is not translated into proteins but plays a critical role in epigenetics. These molecules are called non-coding RNAs (ncRNAs) [47]. They regulate gene expression under physiological and pathological conditions, acting at various stages in protein synthesis at both transcriptional and post-transcriptional levels [80]. Non-coding RNAs are further divided into two main groups, small ncRNAs (shorter than 200 bp) and long ncRNA (lncRNAs; longer than 200 bp).
Out of small ncRNAs, microRNAs (miRs) have been by far the most extensively studied. They are a class of short-chain, linear, approximately 21–25 nucleotides long that negatively regulate gene targets at the post-transcriptional level by perfect complementarity of their “seed” region to 3′-UTR of its target mRNA, inducing their degradation. If there is a mismatch or imperfect complementarity, it results in translational repression [49]. The latest release of the miRbase database (v22) contains 2654 human mature miRs sequences [81], ratifying their importance on gene expression regulation.
Evidence shows that miRs mediate CDDP-induced nephrotoxicity in renal tubular cells through modulation of several molecules involved in a wide array of human (Table 1) and rodents (Table 2) signaling pathways, activating mechanisms such as DNA damage response, oxidative stress, autophagy, necroptosis, cell cycle arrest, apoptosis and inflammatory response (Figure 3). Cisplatin induces apoptosis in renal tubular cells [21,22,23,24,25,26], resulting in p53 activation. However, the epigenetic mechanism regulating cellular death is not well defined. Bhatt et al. demonstrated that induction of miR-34a in vitro during CDDP treatment in BLUMPT cells and in vivo during CDDP-induced nephrotoxicity in C57BL/6 mice were generated by p53. Still, this induction was abrogated in p53-deficient mice [21]. However, other studies showed that CDDP treatment down-regulates miR-122 expression, causing FOXO3 acetylation and activation [36]; simultaneously, miR-34a expression is induced, causing SIRT1 repression [36,37] together with triggering p53 activation [36]. FOXO3 activation also increases BNIP3L expression, a process that is mediated by miR-30c repression [22]. SIRT1 repression is also preceded by miR-449 up-regulation [32]. Upon CDDP-induced damage and consequent p53 activation, BAX is translocated into the mitochondria causing BCL-2 down-regulation, a process mediated by miR-181a [33], miR-449 [32], miR-21-5p and miR-200c [82]; and miR-107 [83]. CDDP treatment has also shown JNK induction [84], which promotes the apoptotic process. The decrease in mitochondrial membrane potential as a consequence of the increased BAX/BCL-2 ratio leads to the release of cytochrome c into the cytosol, binding to other molecules to form the apoptosome and initiating the caspase activation cascade, which is modulated by miRs such as miR-26a [25], -30c [22], -34a [21], -375 [85] and -709 [23].
Although autophagy is generally activated under conditions of nutrient deprivation, it has also been implicated in physiological (development, differentiation) and pathological (neurodegenerative diseases, stress, infection, cancer) processes [86]. Based on this, it has been proven that CDDP also modulates the autophagy process in renal tubular cells [87,88], although the epigenetic mechanisms underlying its occurrence have been poorly studied. However, it is known that CDDP modulates MAPK/ERK1/2 signaling pathway through miR-146b up-regulation [89]. Enhanced mir-146b levels directly impact the expression of mTOR, an autophagy initiation regulatory kinase, which is in turn mediated by miR-199a up-regulation [38] and miR-26b down-regulation [90]. This allows the interaction of Atg13, FIPP200, and ULK1/2 [91], the last one being regulated by miR-141 repression [92]. The Atg13/FIPP200/ULK composite interacts with the PI3K class III complex, composed of several molecules, including Beclin-1, which recruit several ATG proteins, allowing the phagophore formation [91], and thus, starting the autophagy process.
Cisplatin also modulates cellular oxidative stress by regulating NRF2 [93], a transcription factor that regulates the expression of genes encoding antioxidant, anti-inflammatory, and detoxifying proteins [94]. Under normal conditions, NRF2 is located in the cytoplasm, negatively controlled by KEAP1; whereas, under oxidative stress conditions, KEAP1 releases NRF2, which can be translocated to the nucleus and, in this way, binds to antioxidant response elements (ARE), resulting in the transcription of antioxidant genes [95]. Some miRs have been shown to modulate KEAP1/NRF2 interaction, such as miR-192-5p [24] and miR-140-5p [93]. In addition to this, CDDP also alters the redox state, by decreasing the antioxidant activity of GSH, GSH-Px, and SOD [37].
Necroptosis is a programmed form of necrotic cell death considered passive cell death through regulated cell signaling pathways [96]. RIPK1, RIPK3, and MLKL are the main molecules involved in the necroptosis process. Necroptosis is triggered when TNFR1 is activated by TNF-α, resulting in complex I formation. Subsequently, RIPK1, TRADD, and FADD are recruited. Once RIPK1 is activated, its interaction with RIPK3 occurs, forming complex IIb/necrosome, which mediates the phosphorylation of MLKL, whose translocation to the plasma membrane promotes necroptosis by disrupting plasma membrane integrity [97]. Some studies in renal tubular cells have demonstrated the involvement of miR-500a-3p in RIPK1, RIPK3, and MLKL regulation when these cells are treated with CDDP [98,99]. On the other hand, RIPK1 ubiquitination results in NF-κB activation, mediating the survival and production of multiple inflammatory cytokines, which are also regulated by several miRs when these cells are exposed to CDDP. Thus, IL-6 is mediated by miR-494 [15] and miR-34a [37]; IL-8 by miR-500a-3p [99]; and finally, IL-1β by miR-155 [100] and miR-34a [37].
The functional and pathological roles of lncRNAs also have been implicated in CDDP-induced nephrotoxicity. The lncRNA XLOC_032768 can attenuate CDDP-induced apoptosis through TNF-α up-regulation [101]. Similarly, PRNCR1 lncRNA also attenuates CDDP-induced damage in HK-2 cells through up-regulation of EZH1 expression [102]. However, some lncRNAs have been implicated in nephrotoxic damage. Thus, during CDDP-induced damage, lncRNA LRNA9884 increases NF-κB-mediated inflammatory cytokines, especially MIF [103]. Likewise, lncRNA GAS5 can increase CDDP-induced apoptosis of HK-2 cells by modulating miR-205-5p [104].
Circular RNAs (circRNAs) are single-stranded, covalently locked transcripts produced from a precursor mRNA [105]. These circRNAs can act as sponges for miRs, resulting in increased expression [49]. In this way, circ-0114427 attracts miR-494 and interacts with it, increasing ATF3 expression and reducing inflammatory cytokines, thus exerting a nephroprotective role against CDDP damage [15].

6. Potential Challenges and Future Directions

Epigenetics is an active research field concerning cisplatin-induced nephrotoxicity. So far, studies show that the mechanisms involved play a significant role in the pathogenesis of this condition. However, little is known about the interaction between different epigenetic marks. Nonetheless, crosstalk between epigenetic modifications does exist in this pathology. For example, during CDDP-induced nephrotoxicity, lncRNA LRNA9884 increases NF-κB expression [103]. On the other hand, miR-34 [21,36], miR-192-5p [24] and, miR-449 [32] activate p53. Both NF-κB and p53 up-regulate miR-375 [85], inducing apoptosis in renal proximal tubular cells. However, inhibiting p300 attenuates this adverse effect due to reduced acetylation of NF-κB [61]. In the same way, overexpression of SIRT1 and SIRT6 inhibits p53 signaling, reducing cisplatin-induced damage [43,74]. Clearly, there are many gaps regarding epigenetic modulation in the pathogenesis of this condition, especially in the cross-talking for these mechanisms. Understanding the way that DNA methylation, histone modifications, and ncRNAs underlie this disease will provide us with the opportunity to discover new biomarkers for early diagnosis, as well as the establishment of novel therapies or treatments aiming to decrease the incidence of cisplatin-induced nephrotoxicity.

7. Conclusions

Nephrotoxicity is the most lethal side effect of cisplatin treatment. During the last few years, rapid progress has been made in understanding the contribution of epigenetic mechanisms underlying nephrotoxicity to be used as potential biomarkers for early diagnosis and/or to employ different therapeutic strategies. These studies demonstrate the complexity of the interactions of epigenetic modifications, which have provided a better understanding of the epigenetic regulation of CDDP-induced nephrotoxicity. Continued research in this area will provide new therapeutic targets and early, sensitive, and specific diagnostic biomarkers to detect CDDP-induced nephrotoxicity.

Author Contributions

Conceptualization, P.L. and L.A.S.; methodology, P.L.; software, P.L.; validation, P.L., K.S., N.S. and P.M.; formal analysis, P.L. and N.S.; investigation, P.L.; resources, L.A.S.; data curation, P.L.; writing—original draft preparation, P.L.; writing—review and editing, P.L., T.Z., P.M. and L.A.S.; visualization, L.A.S.; supervision, L.A.S.; project administration, P.M. and L.A.S.; funding acquisition, N.S., K.S., P.M. and L.A.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by ANID-FAPESP (Grant Number 19/13250-1), FONDECYT (Grant Number 1171765) and Programa de Formación de Investigadores Postdoctorales, VRIP, Universidad de La Frontera, Temuco, Chile (Code VRIP 21001).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. WHO. World Health Statistics 2020: Monitoring Health for the SDGs; World Health Organization: Geneva, Switzerland, 2020; Volume 21, pp. 1–9. [Google Scholar] [CrossRef]
  2. Vermorken, J.B.; Remenar, E.; van Herpen, C.; Gorlia, T.; Mesia, R.; Degardin, M.; Stewart, J.S.; Jelic, S.; Betka, J.; Preiss, J.H.; et al. Cisplatin, fluorouracil, and docetaxel in unresectable head and neck cancer. N. Engl. J. Med. 2007, 357, 1695–1704. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Motzer, R.J.; Sheinfeld, J.; Mazumdar, M.; Bajorin, D.F.; Bosl, G.J.; Herr, H.; Lyn, P.; Vlamis, V. Etoposide and cisplatin adjuvant therapy for patients with pathologic stage II germ cell tumors. J. Clin. Oncol. 1995, 13, 2700–2704. [Google Scholar] [CrossRef]
  4. Li, D.; Zhang, Y.; Xie, Y.; Xiang, J.; Zhu, Y.; Yang, J. Enhanced tumor suppression by adenoviral PTEN gene therapy combined with cisplatin chemotherapy in small-cell lung cancer. Cancer Gene Ther. 2013, 20, 251–259. [Google Scholar] [CrossRef] [PubMed]
  5. Magali, L.; Pascal, F.; Serge, A.; Mathieu, B.; Ayoube, Z.; Claire, T.; Christiane, M. Better survival in impaired renal function patients with metastatic non-small cell lung cancer treated by cisplatin-pemetrexed. Eur. J. Clin. Pharmacol. 2020, 76, 1573–1580. [Google Scholar] [CrossRef]
  6. Armstrong, D.K.; Bundy, B.; Wenzel, L.; Huang, H.Q.; Baergen, R.; Lele, S.; Copeland, L.J.; Walker, J.L.; Burger, R.A. Intraperitoneal cisplatin and paclitaxel in ovarian cancer. N. Engl. J. Med. 2006, 354, 34–43. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Moore, K.N.; Herzog, T.J.; Lewin, S.; Giuntoli, R.L.; Armstrong, D.K.; Rocconi, R.P.; Spannuth, W.A.; Gold, M.A. A comparison of cisplatin/paclitaxel and carboplatin/paclitaxel in stage IVB, recurrent or persistent cervical cancer. Gynecol. Oncol. 2007, 105, 299–303. [Google Scholar] [CrossRef]
  8. Coppin, C.M.; Gospodarowicz, M.K.; James, K.; Tannock, I.F.; Zee, B.; Carson, J.; Pater, J.; Sullivan, L.D. Improved local control of invasive bladder cancer by concurrent cisplatin and preoperative or definitive radiation. The National Cancer Institute of Canada Clinical Trials Group. J. Clin. Oncol. 1996, 14, 2901–2907. [Google Scholar] [CrossRef]
  9. Miller, R.P.; Tadagavadi, R.K.; Ramesh, G.; Reeves, W.B. Mechanisms of Cisplatin nephrotoxicity. Toxins 2010, 2, 2490–2518. [Google Scholar] [CrossRef] [Green Version]
  10. Arany, I.; Safirstein, R.L. Cisplatin nephrotoxicity. Semin. Nephrol. 2003, 23, 460–464. [Google Scholar] [CrossRef]
  11. Al-Naimi, M.; Rasheed, H.; Hussien, N.; Al-Kuraishy, H.; Al-Gareeb, A. Nephrotoxicity: Role and significance of renal biomarkers in the early detection of acute renal injury. J. Adv. Pharm. Technol. Res. 2019, 10, 95–99. [Google Scholar] [CrossRef]
  12. Griffin, B.R.; Faubel, S.; Edelstein, C.L. Biomarkers of drug-induced kidney toxicity. Ther. Drug Monit. 2019, 41, 213–226. [Google Scholar] [CrossRef]
  13. Alfieri, A.B.; Cubeddu, L.X. Role of NK1 receptors on cisplatin-induced nephrotoxicity in the rat. Naunyn Schmiedebergs Arch. Pharmacol. 2000, 361, 334–338. [Google Scholar] [CrossRef]
  14. Zhu, S.; Pabla, N.; Tang, C.; He, L.; Dong, Z. DNA damage response in cisplatin-induced nephrotoxicity. Arch. Toxicol. 2015, 89, 2197–2205. [Google Scholar] [CrossRef] [Green Version]
  15. Cao, Y.; Mi, X.; Zhang, D.; Wang, Z.; Zuo, Y.; Tang, W. Transcriptome sequencing of circular RNA reveals a novel circular RNA-has_circ_0114427 in the regulation of inflammation in acute kidney injury. Clin. Sci. 2020, 134, 139–154. [Google Scholar] [CrossRef] [PubMed]
  16. Wu, C.; Morris, J.R. Genes, genetics, and epigenetics: A correspondence. Science 2001, 293, 1103–1105. [Google Scholar] [CrossRef] [Green Version]
  17. Calcagno, D.Q.; Gigek, C.O.; Chen, E.S.; Burbano, R.R.; Smith Mde, A. DNA and histone methylation in gastric carcinogenesis. World J. Gastroenterol. 2013, 19, 1182–1192. [Google Scholar] [CrossRef] [PubMed]
  18. Lau, J.K.; Ensing, B. Hydrolysis of cisplatin--a first-principles metadynamics study. Phys. Chem. Chem. Phys. 2010, 12, 10348–10355. [Google Scholar] [CrossRef]
  19. Galgamuwa, R. An Investigation Into The Prevention Of Cisplatin-Induced Nephrotoxicity By Dichloroacetate. Australian National University. Available online: https://openresearch-repository.anu.edu.au/ (accessed on 1 May 2021).
  20. Pabla, N.; Huang, S.; Mi, Q.S.; Daniel, R.; Dong, Z. ATR-Chk2 signaling in p53 activation and DNA damage response during cisplatin-induced apoptosis. J. Biol. Chem. 2008, 283, 6572–6583. [Google Scholar] [CrossRef] [Green Version]
  21. Bhatt, K.; Zhou, L.; Mi, Q.S.; Huang, S.; She, J.X.; Dong, Z. MicroRNA-34a is induced via p53 during cisplatin nephrotoxicity and contributes to cell survival. Mol. Med. 2010, 16, 409–416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Du, B.; Dai, X.M.; Li, S.; Qi, G.L.; Cao, G.X.; Zhong, Y.; Yin, P.D.; Yang, X.S. MiR-30c regulates cisplatin-induced apoptosis of renal tubular epithelial cells by targeting Bnip3L and Hspa5. Cell Death Dis. 2017, 8, e2987. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Guo, Y.; Ni, J.; Chen, S.; Bai, M.; Lin, J.; Ding, G.; Zhang, Y.; Sun, P.; Jia, Z.; Huang, S.; et al. MicroRNA-709 Mediates Acute Tubular Injury through Effects on Mitochondrial Function. J. Am. Soc. Nephrol. 2018, 29, 449–461. [Google Scholar] [CrossRef] [Green Version]
  24. Zou, D.; Ganugula, R.; Arora, M.; Nabity, M.B.; Sheikh-Hamad, D.; Kumar, M. Oral delivery of nanoparticle urolithin A normalizes cellular stress and improves survival in mouse model of cisplatin-induced AKI. Am. J. Physiol. Renal Physiol. 2019, 317, F1255–F1264. [Google Scholar] [CrossRef]
  25. Yang, Y.; Li, Z.L.; Wang, F.M.; Tang, R.N.; Tu, Y.; Liu, H. MicroRNA26a inhibits cisplatin-induced renal tubular epithelial cells apoptosis through suppressing the expression of transient receptor potential channel 6 mediated dynamin-related protein 1. Cell Biochem. Funct. 2020, 38, 384–391. [Google Scholar] [CrossRef]
  26. Harrill, A.H.; Lin, H.; Tobacyk, J.; Seely, J.C. Mouse population-based evaluation of urinary protein and miRNA biomarker performance associated with cisplatin renal injury. Exp. Biol. Med. 2018, 243, 237–247. [Google Scholar] [CrossRef]
  27. Liu, Z.; Li, Z.; Chen, Z.; Li, C.; Lei, L.; Wu, X.; Li, Y. Numb ameliorates necrosis and inflammation in acute kidney injury induced by cisplatin. Chem. Biol. Interact. 2020, 330, 109251. [Google Scholar] [CrossRef]
  28. Ramesh, G.; Reeves, W.B. TNFR2-mediated apoptosis and necrosis in cisplatin-induced acute renal failure. Am. J. Physiol. Renal Physiol. 2003, 285, F610–F618. [Google Scholar] [CrossRef]
  29. Landau, S.I.; Guo, X.; Velazquez, H.; Torres, R.; Olson, E.; Garcia-Milian, R.; Moeckel, G.W.; Desir, G.V.; Safirstein, R. Regulated necrosis and failed repair in cisplatin-induced chronic kidney disease. Kidney Int. 2019, 95, 797–814. [Google Scholar] [CrossRef] [PubMed]
  30. Nosaka, K.; Nakada, J.; Endou, H. Cisplatin-induced alterations in renal structure, ammoniagenesis and gluconeogenesis of rats. Kidney Int. 1992, 41, 73–79. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Qian, W.; Nishikawa, M.; Haque, A.M.; Hirose, M.; Mashimo, M.; Sato, E.; Inoue, M. Mitochondrial density determines the cellular sensitivity to cisplatin-induced cell death. Am. J. Physiol. Cell Physiol. 2005, 289, C1466–C1475. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Qin, W.; Xie, W.; Yang, X.; Xia, N.; Yang, K. Inhibiting microRNA-449 Attenuates Cisplatin-Induced Injury in NRK-52E Cells Possibly via Regulating the SIRT1/P53/BAX Pathway. Med. Sci. Monit. 2016, 22, 818–823. [Google Scholar] [CrossRef]
  33. Zhu, H.Y.; Liu, M.Y.; Hong, Q.; Zhang, D.; Geng, W.J.; Xie, Y.S.; Chen, X.M. Role of microRNA-181a in the apoptosis of tubular epithelial cell induced by cisplatin. Chin. Med. J. 2012, 125, 523–526. [Google Scholar] [CrossRef]
  34. McDonald, E.S.; Windebank, A.J. Cisplatin-induced apoptosis of DRG neurons involves bax redistribution and cytochrome c release but not fas receptor signaling. Neurobiol. Dis. 2002, 9, 220–233. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Ni, J.; Hou, X.; Wang, X.; Shi, Y.; Xu, L.; Zheng, X.; Liu, N.; Qiu, A.; Zhuang, S. 3-deazaneplanocin A protects against cisplatin-induced renal tubular cell apoptosis and acute kidney injury by restoration of E-cadherin expression. Cell Death Dis. 2019, 10, 355. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Lee, C.G.; Kim, J.G.; Kim, H.J.; Kwon, H.K.; Cho, I.J.; Choi, D.W.; Lee, W.H.; Kim, W.D.; Hwang, S.J.; Choi, S.; et al. Discovery of an integrative network of microRNAs and transcriptomics changes for acute kidney injury. Kidney Int. 2014, 86, 943–953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Zhang, Y.; Tao, X.; Yin, L.; Xu, L.; Xu, Y.; Qi, Y.; Han, X.; Song, S.; Zhao, Y.; Lin, Y.; et al. Protective effects of dioscin against cisplatin-induced nephrotoxicity via the microRNA-34a/sirtuin 1 signalling pathway. Br. J. Pharmacol. 2017, 174, 2512–2527. [Google Scholar] [CrossRef] [Green Version]
  38. Yang, A.; Liu, F.; Guan, B.; Luo, Z.; Lin, J.; Fang, W.; Liu, L.; Zuo, W. p53 induces miR-199a-3p to suppress mechanistic target of rapamycin activation in cisplatin-induced acute kidney injury. J. Cell Biochem. 2019, 120, 17625–17634. [Google Scholar] [CrossRef] [PubMed]
  39. Suter-Dick, L.; Mauch, L.; Ramp, D.; Caj, M.; Vormann, M.K.; Hutter, S.; Lanz, H.L.; Vriend, J.; Masereeuw, R.; Wilmer, M.J. Combining Extracellular miRNA Determination with Microfluidic 3D Cell Cultures for the Assessment of Nephrotoxicity: A Proof of Concept Study. AAPS J. 2018, 20, 86. [Google Scholar] [CrossRef]
  40. Yang, C.; Kaushal, V.; Haun, R.S.; Seth, R.; Shah, S.V.; Kaushal, G.P. Transcriptional activation of caspase-6 and -7 genes by cisplatin-induced p53 and its functional significance in cisplatin nephrotoxicity. Cell Death Differ. 2008, 15, 530–544. [Google Scholar] [CrossRef]
  41. Sun, L.; Liu, J.; Yuan, Y.; Zhang, X.; Dong, Z. Protective effect of the BET protein inhibitor JQ1 in cisplatin-induced nephrotoxicity. Am. J. Physiol. Renal Physiol. 2018, 315, F469–F478. [Google Scholar] [CrossRef]
  42. Sun, C.Y.; Nie, J.; Zheng, Z.L.; Zhao, J.; Wu, L.M.; Zhu, Y.; Su, Z.Q.; Zheng, G.J.; Feng, B. Renoprotective effect of scutellarin on cisplatin-induced renal injury in mice: Impact on inflammation, apoptosis, and autophagy. Biomed. Pharmacother. 2019, 112, 108647. [Google Scholar] [CrossRef]
  43. Li, Z.; Xu, K.; Zhang, N.; Amador, G.; Wang, Y.; Zhao, S.; Li, L.; Qiu, Y.; Wang, Z. Overexpressed SIRT6 attenuates cisplatin-induced acute kidney injury by inhibiting ERK1/2 signaling. Kidney Int. 2018, 93, 881–892. [Google Scholar] [CrossRef]
  44. Almaghrabi, O.A. Molecular and biochemical investigations on the effect of quercetin on oxidative stress induced by cisplatin in rat kidney. Saudi J. Biol. Sci. 2015, 22, 227–231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Kruidering, M.; Van de Water, B.; de Heer, E.; Mulder, G.J.; Nagelkerke, J.F. Cisplatin-induced nephrotoxicity in porcine proximal tubular cells: Mitochondrial dysfunction by inhibition of complexes I to IV of the respiratory chain. J. Pharmacol. Exp. Ther. 1997, 280, 638–649. [Google Scholar]
  46. Hao, Y.; Guo, F.; Huang, Z.; Feng, Y.; Xia, Z.; Liu, J.; Li, L.; Huang, R.; Lin, L.; Ma, L.; et al. 2-Methylquinazoline derivative 23BB as a highly selective histone deacetylase 6 inhibitor alleviated cisplatin-induced acute kidney injury. Biosci. Rep. 2020, 40. [Google Scholar] [CrossRef]
  47. Xiang, X.; Guo, C.; Tang, C.; Cai, J.; Dong, Z. Epigenetic Regulation in Kidney Toxicity: Insights From Cisplatin Nephrotoxicity. Semin. Nephrol. 2019, 39, 152–158. [Google Scholar] [CrossRef] [PubMed]
  48. Saldanha, S.N.; Tollefsbol, T.O. Epigenetic Approaches to Cancer Therapy. In Epigenetics in Human Disease; Tollefsbol, T.O., Ed.; Academic Press: Cambridge, MA, USA, 2018; Volume 6, pp. 219–247. [Google Scholar]
  49. Guo, C.; Dong, G.; Liang, X.; Dong, Z. Epigenetic regulation in AKI and kidney repair: Mechanisms and therapeutic implications. Nat. Rev. Nephrol. 2019, 15, 220–239. [Google Scholar] [CrossRef] [PubMed]
  50. Guo, C.; Pei, L.; Xiao, X.; Wei, Q.; Chen, J.K.; Ding, H.F.; Huang, S.; Fan, G.; Shi, H.; Dong, Z. DNA methylation protects against cisplatin-induced kidney injury by regulating specific genes, including interferon regulatory factor 8. Kidney Int. 2017, 92, 1194–1205. [Google Scholar] [CrossRef] [PubMed]
  51. Scholpa, N.E.; Kolli, R.T.; Moore, M.; Arnold, R.D.; Glenn, T.C.; Cummings, B.S. Nephrotoxicity of epigenetic inhibitors used for the treatment of cancer. Chem. Biol. Interact. 2016, 258, 21–29. [Google Scholar] [CrossRef] [Green Version]
  52. Tikoo, K.; Ali, I.Y.; Gupta, J.; Gupta, C. 5-Azacytidine prevents cisplatin induced nephrotoxicity and potentiates anticancer activity of cisplatin by involving inhibition of metallothionein, pAKT and DNMT1 expression in chemical induced cancer rats. Toxicol. Lett. 2009, 191, 158–166. [Google Scholar] [CrossRef]
  53. Qian, C.Y.; Zheng, Y.; Wang, Y.; Chen, J.; Liu, J.Y.; Zhou, H.H.; Yin, J.Y.; Liu, Z.Q. Associations of genetic polymorphisms of the transporters organic cation transporter 2 (OCT2), multidrug and toxin extrusion 1 (MATE1), and ATP-binding cassette subfamily C member 2 (ABCC2) with platinum-based chemotherapy response and toxicity in non-small cell lung cancer patients. Chin. J. Cancer 2016, 35, 85. [Google Scholar] [CrossRef] [Green Version]
  54. Aoki, M.; Terada, T.; Kajiwara, M.; Ogasawara, K.; Ikai, I.; Ogawa, O.; Katsura, T.; Inui, K. Kidney-specific expression of human organic cation transporter 2 (OCT2/SLC22A2) is regulated by DNA methylation. Am. J. Physiol. Renal Physiol. 2008, 295, F165–F170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Mercurio, C.; Plyte, S.; Minucci, S. Chapter 4—Alterations of Histone Modifications in Cancer. In Epigenetics in Human Disease; Translational Epigenetics; Elsevier Inc.: Amsterdam, The Netherlands, 2012; pp. 53–87. [Google Scholar] [CrossRef]
  56. Cheng, L.; Li, C.; Xi, Z.; Wei, K.; Yuan, S.; Arnesano, F.; Natile, G.; Liu, Y. Cisplatin reacts with histone H1 and the adduct forms a ternary complex with DNA. Metallomics 2019, 11, 556–564. [Google Scholar] [CrossRef] [PubMed]
  57. Chen, S.Q.; Li, J.Q.; Wang, X.Q.; Lei, W.J.; Li, H.; Wan, J.; Hu, Z.; Zou, Y.W.; Wu, X.Y.; Niu, H.X. EZH2-inhibitor DZNep enhances apoptosis of renal tubular epithelial cells in presence and absence of cisplatin. Cell Div. 2020, 15, 8. [Google Scholar] [CrossRef] [PubMed]
  58. Hosseini, A.; Minucci, S. Chapter 6—Alterations of Histone Modifications in Cancer. In Epigenetics in Human Disease; Tollefsbol, T.O., Ed.; Academic Press: Cambridge, MA, USA; Elsevier Inc.: Amsterdam, The Netherlands, 2018; Volume 6, pp. 141–217. [Google Scholar]
  59. Gu, W.; Roeder, R.G. Activation of p53 sequence-specific DNA binding by acetylation of the p53 C-terminal domain. Cell 1997, 90, 595–606. [Google Scholar] [CrossRef] [Green Version]
  60. Chen, L.F.; Mu, Y.; Greene, W.C. Acetylation of RelA at discrete sites regulates distinct nuclear functions of NF-kappaB. EMBO J. 2002, 21, 6539–6548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Kim, J.Y.; Jo, J.; Leem, J.; Park, K.K. Inhibition of p300 by Garcinol Protects against Cisplatin-Induced Acute Kidney Injury through Suppression of Oxidative Stress, Inflammation, and Tubular Cell Death in Mice. Antioxidants 2020, 9, 1271. [Google Scholar] [CrossRef]
  62. Jiang, W.; Yuan, X.; Zhu, H.; He, C.; Ge, C.; Tang, Q.; Xu, C.; Hu, B.; Huang, C.; Ma, T. Inhibition of Histone H3K27 Acetylation Orchestrates Interleukin-9-Mediated and Plays an Anti-Inflammatory Role in Cisplatin-Induced Acute Kidney Injury. Front. Immunol. 2020, 11, 231. [Google Scholar] [CrossRef] [Green Version]
  63. Ma, T.; Huang, C.; Xu, Q.; Yang, Y.; Liu, Y.; Meng, X.; Li, J.; Ye, M.; Liang, H. Suppression of BMP-7 by histone deacetylase 2 promoted apoptosis of renal tubular epithelial cells in acute kidney injury. Cell Death Dis. 2017, 8, e3139. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Sakao, Y.; Kato, A.; Tsuji, T.; Yasuda, H.; Togawa, A.; Fujigaki, Y.; Kahyo, T.; Setou, M.; Hishida, A. Cisplatin induces Sirt1 in association with histone deacetylation and increased Werner syndrome protein in the kidney. Clin. Exp. Nephrol. 2011, 15, 363–372. [Google Scholar] [CrossRef] [Green Version]
  65. Tang, J.; Shi, Y.; Liu, N.; Xu, L.; Zang, X.; Li, P.; Zhang, J.; Zheng, X.; Qiu, A.; Zhuang, S. Blockade of histone deacetylase 6 protects against cisplatin-induced acute kidney injury. Clin. Sci 2018, 132, 339–359. [Google Scholar] [CrossRef]
  66. Mikami, D.; Kobayashi, M.; Uwada, J.; Yazawa, T.; Kamiyama, K.; Nishimori, K.; Nishikawa, Y.; Morikawa, Y.; Yokoi, S.; Takahashi, N.; et al. beta-Hydroxybutyrate, a ketone body, reduces the cytotoxic effect of cisplatin via activation of HDAC5 in human renal cortical epithelial cells. Life Sci. 2019, 222, 125–132. [Google Scholar] [CrossRef] [PubMed]
  67. Dong, G.; Luo, J.; Kumar, V.; Dong, Z. Inhibitors of histone deacetylases suppress cisplatin-induced p53 activation and apoptosis in renal tubular cells. Am. J. Physiol. Renal Physiol. 2010, 298, F293–F300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Tang, J.; Zhuang, S. Upregulation of AMWAP: A novel mechanism for HDAC inhibitors to protect against cisplatin nephrotoxicity. Kidney Int. 2016, 89, 267–269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Ranganathan, P.; Hamad, R.; Mohamed, R.; Jayakumar, C.; Muthusamy, T.; Ramesh, G. Histone deacetylase-mediated silencing of AMWAP expression contributes to cisplatin nephrotoxicity. Kidney Int. 2016, 89, 317–326. [Google Scholar] [CrossRef] [Green Version]
  70. Zhu, H.; Jiang, W.; Zhao, H.; He, C.; Tang, X.; Xu, S.; Xu, C.; Feng, R.; Li, J.; Ma, T.; et al. PSTPIP2 inhibits cisplatin-induced acute kidney injury by suppressing apoptosis of renal tubular epithelial cells. Cell Death Dis. 2020, 11, 1057. [Google Scholar] [CrossRef]
  71. Arany, I.; Herbert, J.; Herbert, Z.; Safirstein, R.L. Restoration of CREB function ameliorates cisplatin cytotoxicity in renal tubular cells. Am. J. Physiol. Renal Physiol. 2008, 294, F577–F581. [Google Scholar] [CrossRef] [Green Version]
  72. Liu, J.; Livingston, M.J.; Dong, G.; Tang, C.; Su, Y.; Wu, G.; Yin, X.M.; Dong, Z. Histone deacetylase inhibitors protect against cisplatin-induced acute kidney injury by activating autophagy in proximal tubular cells. Cell Death Dis. 2018, 9, 322. [Google Scholar] [CrossRef]
  73. Jung, Y.J.; Lee, J.E.; Lee, A.S.; Kang, K.P.; Lee, S.; Park, S.K.; Lee, S.Y.; Han, M.K.; Kim, D.H.; Kim, W. SIRT1 overexpression decreases cisplatin-induced acetylation of NF-kappaB p65 subunit and cytotoxicity in renal proximal tubule cells. Biochem. Biophys. Res. Commun. 2012, 419, 206–210. [Google Scholar] [CrossRef]
  74. Kim, D.H.; Jung, Y.J.; Lee, J.E.; Lee, A.S.; Kang, K.P.; Lee, S.; Park, S.K.; Han, M.K.; Lee, S.Y.; Ramkumar, K.M.; et al. SIRT1 activation by resveratrol ameliorates cisplatin-induced renal injury through deacetylation of p53. Am. J. Physiol. Renal Physiol. 2011, 301, F427–F435. [Google Scholar] [CrossRef] [Green Version]
  75. Kim, J.Y.; Jo, J.; Kim, K.; An, H.J.; Gwon, M.G.; Gu, H.; Kim, H.J.; Yang, A.Y.; Kim, S.W.; Jeon, E.J.; et al. Pharmacological Activation of Sirt1 Ameliorates Cisplatin-Induced Acute Kidney Injury by Suppressing Apoptosis, Oxidative Stress, and Inflammation in Mice. Antioxidants 2019, 8, 322. [Google Scholar] [CrossRef] [Green Version]
  76. Jung, Y.J.; Park, W.; Kang, K.P.; Kim, W. SIRT2 is involved in cisplatin-induced acute kidney injury through regulation of mitogen-activated protein kinase phosphatase-1. Nephrol. Dial. Transplant. 2020, 35, 1145–1156. [Google Scholar] [CrossRef] [Green Version]
  77. Yoon, S.P.; Kim, J. Poly(ADP-ribose) polymerase 1 contributes to oxidative stress through downregulation of sirtuin 3 during cisplatin nephrotoxicity. Anat. Cell Biol. 2016, 49, 165–176. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Kim, D.; Park, W.; Lee, S.; Kim, W.; Park, S.K.; Kang, K.P. Absence of Sirt3 aggravates cisplatin nephrotoxicity via enhanced renal tubular apoptosis and inflammation. Mol. Med. Rep. 2018, 18, 3665–3672. [Google Scholar] [CrossRef] [Green Version]
  79. Chiba, T.; Peasley, K.D.; Cargill, K.R.; Maringer, K.V.; Bharathi, S.S.; Mukherjee, E.; Zhang, Y.; Holtz, A.; Basisty, N.; Yagobian, S.D.; et al. Sirtuin 5 Regulates Proximal Tubule Fatty Acid Oxidation to Protect against AKI. J. Am. Soc. Nephrol. 2019, 30, 2384–2398. [Google Scholar] [CrossRef] [PubMed]
  80. Bertrand-Lehouillier, V.; Legault, L.-M.; McGraw, S. Endocrine Epigenetics, Epigenetic Profiling and Biomarker Identification. In Encyclopedia of Endocrine Diseases; Huhtaniemi, I., Martini, L., Eds.; Academic Press: Cambridge, MA, USA, 2019; pp. 31–35. [Google Scholar] [CrossRef]
  81. Kozomara, A.; Birgaoanu, M.; Griffiths-Jones, S. miRBase: From microRNA sequences to function. Nucleic Acids Res. 2019, 47, D155–D162. [Google Scholar] [CrossRef] [PubMed]
  82. Pavkovic, M.; Robinson-Cohen, C.; Chua, A.S.; Nicoara, O.; Cardenas-Gonzalez, M.; Bijol, V.; Ramachandran, K.; Hampson, L.; Pirmohamed, M.; Antoine, D.J.; et al. Detection of Drug-Induced Acute Kidney Injury in Humans Using Urinary KIM-1, miR-21, -200c, and -423. Toxicol. Sci. 2016, 152, 205–213. [Google Scholar] [CrossRef] [PubMed]
  83. Sun, W.; Zhu, Q.; Yan, L.; Shao, F. Mesenchymal stem cells alleviate acute kidney injury via miR-107-mediated regulation of ribosomal protein S19. Ann. Transl. Med. 2019, 7, 765. [Google Scholar] [CrossRef]
  84. Wu, J.; Li, D.D.; Li, J.Y.; Yin, Y.C.; Li, P.C.; Qiu, L.; Chen, L.M. Identification of microRNA-mRNA networks involved in cisplatin-induced renal tubular epithelial cells injury. Eur. J. Pharmacol. 2019, 851, 1–12. [Google Scholar] [CrossRef]
  85. Hao, J.; Lou, Q.; Wei, Q.; Mei, S.; Li, L.; Wu, G.; Mi, Q.S.; Mei, C.; Dong, Z. MicroRNA-375 Is Induced in Cisplatin Nephrotoxicity to Repress Hepatocyte Nuclear Factor 1-beta. J. Biol. Chem. 2017, 292, 4571–4582. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Khandia, R.; Dadar, M.; Munjal, A.; Dhama, K.; Karthik, K.; Tiwari, R.; Yatoo, M.I.; Iqbal, H.M.N.; Singh, K.P.; Joshi, S.K.; et al. A Comprehensive Review of Autophagy and Its Various Roles in Infectious, Non-Infectious, and Lifestyle Diseases: Current Knowledge and Prospects for Disease Prevention, Novel Drug Design, and Therapy. Cells 2019, 8, 674. [Google Scholar] [CrossRef] [Green Version]
  87. Yang, C.; Kaushal, V.; Shah, S.V.; Kaushal, G.P. Autophagy is associated with apoptosis in cisplatin injury to renal tubular epithelial cells. Am. J. Physiol. Renal Physiol. 2008, 294, F777–F787. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Takahashi, A.; Kimura, T.; Takabatake, Y.; Namba, T.; Kaimori, J.; Kitamura, H.; Matsui, I.; Niimura, F.; Matsusaka, T.; Fujita, N.; et al. Autophagy guards against cisplatin-induced acute kidney injury. Am. J. Pathol. 2012, 180, 517–525. [Google Scholar] [CrossRef] [PubMed]
  89. Zhu, Y.; Yu, J.; Yin, L.; Zhou, Y.; Sun, Z.; Jia, H.; Tao, Y.; Liu, W.; Zhang, B.; Zhang, J.; et al. MicroRNA-146b, a Sensitive Indicator of Mesenchymal Stem Cell Repair of Acute Renal Injury. Stem Cells Transl. Med. 2016, 5, 1406–1415. [Google Scholar] [CrossRef] [Green Version]
  90. El Magdoub, H.M.; Schaalan, M.F.; Rahmo, R.M.; Farag, D.B.; Khedr, L.H. Implications of miRNAs on TGF-beta/TAK1/mTOR pathway in mediating the renoprotective effects of pentoxifylline against cisplatin-induced nephrotoxicity in rats. Toxicol. Appl. Pharmacol. 2020, 404, 115184. [Google Scholar] [CrossRef]
  91. Jung, C.H.; Ro, S.H.; Cao, J.; Otto, N.M.; Kim, D.H. mTOR regulation of autophagy. FEBS Lett. 2010, 584, 1287–1295. [Google Scholar] [CrossRef] [Green Version]
  92. De Almeida, D.C.; Bassi, E.J.; Azevedo, H.; Anderson, L.; Origassa, C.S.; Cenedeze, M.A.; de Andrade-Oliveira, V.; Felizardo, R.J.; da Silva, R.C.; Hiyane, M.I.; et al. A Regulatory miRNA-mRNA Network Is Associated with Tissue Repair Induced by Mesenchymal Stromal Cells in Acute Kidney Injury. Front. Immunol. 2016, 7, 645. [Google Scholar] [CrossRef] [Green Version]
  93. Liao, W.; Fu, Z.; Zou, Y.; Wen, D.; Ma, H.; Zhou, F.; Chen, Y.; Zhang, M.; Zhang, W. MicroRNA-140-5p attenuated oxidative stress in Cisplatin induced acute kidney injury by activating Nrf2/ARE pathway through a Keap1-independent mechanism. Exp. Cell Res. 2017, 360, 292–302. [Google Scholar] [CrossRef]
  94. Loboda, A.; Damulewicz, M.; Pyza, E.; Jozkowicz, A.; Dulak, J. Role of Nrf2/HO-1 system in development, oxidative stress response and diseases: An evolutionarily conserved mechanism. Cell Mol. Life Sci. 2016, 73, 3221–3247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Lu, M.C.; Ji, J.A.; Jiang, Z.Y.; You, Q.D. The Keap1-Nrf2-ARE Pathway As a Potential Preventive and Therapeutic Target: An Update. Med. Res. Rev. 2016, 36, 924–963. [Google Scholar] [CrossRef]
  96. Xu, Y.; Ma, H.; Shao, J.; Wu, J.; Zhou, L.; Zhang, Z.; Wang, Y.; Huang, Z.; Ren, J.; Liu, S.; et al. A Role for Tubular Necroptosis in Cisplatin-Induced AKI. J. Am. Soc. Nephrol. 2015, 26, 2647–2658. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Najafov, A.; Chen, H.; Yuan, J. Necroptosis and Cancer. Trends Cancer 2017, 3, 294–301. [Google Scholar] [CrossRef] [Green Version]
  98. Zhang, S.; Sun, H.; Kong, W.; Zhang, B. Functional role of microRNA-500a-3P-loaded liposomes in the treatment of cisplatin-induced AKI. IET Nanobiotechnol. 2020, 14, 465–469. [Google Scholar] [CrossRef] [PubMed]
  99. Jiang, L.; Liu, X.Q.; Ma, Q.; Yang, Q.; Gao, L.; Li, H.D.; Wang, J.N.; Wei, B.; Wen, J.; Li, J.; et al. hsa-miR-500a-3P alleviates kidney injury by targeting MLKL-mediated necroptosis in renal epithelial cells. FASEB J. 2019, 33, 3523–3535. [Google Scholar] [CrossRef] [Green Version]
  100. Pellegrini, K.L.; Han, T.; Bijol, V.; Saikumar, J.; Craciun, F.L.; Chen, W.W.; Fuscoe, J.C.; Vaidya, V.S. MicroRNA-155 deficient mice experience heightened kidney toxicity when dosed with cisplatin. Toxicol. Sci. 2014, 141, 484–492. [Google Scholar] [CrossRef] [Green Version]
  101. Zhou, X.; Jiang, K.; Luo, H.; Wu, C.; Yu, W.; Cheng, F. Novel lncRNA XLOC_032768 alleviates cisplatin-induced apoptosis and inflammatory response of renal tubular epithelial cells through TNF-alpha. Int. Immunopharmacol. 2020, 83, 106472. [Google Scholar] [CrossRef] [PubMed]
  102. Li, J.; Fan, X.; Wang, Q.; Gong, Y.; Guo, L. Long Noncoding RNA PRNCR1 Reduces Renal Epithelial Cell Apoptosis in Cisplatin-Induced AKI by Regulating miR-182-5p/EZH1. Kidney Blood Press. Res. 2021, 46, 162–172. [Google Scholar] [CrossRef]
  103. Zhang, Y.; Tang, P.M.; Niu, Y.; Garcia Cordoba, C.A.; Huang, X.R.; Yu, C.; Lan, H.Y. Long Non-coding RNA LRNA9884 Promotes Acute Kidney Injury via Regulating NF-kB-Mediated Transcriptional Activation of MIF. Front. Physiol. 2020, 11, 590027. [Google Scholar] [CrossRef] [PubMed]
  104. Zhang, M.; Bi, H.; Wang, S.; Sun, X.; Li, Y. Long non-coding RNA GAS5 aggravate renal epithelial cell apoptosis in cisplatin-induced AKI by regulating miR-205-5p. Res. Sq. 2021. [Google Scholar] [CrossRef]
  105. Huang, S.; Yang, B.; Chen, B.J.; Bliim, N.; Ueberham, U.; Arendt, T.; Janitz, M. The emerging role of circular RNAs in transcriptome regulation. Genomics 2017, 109, 401–407. [Google Scholar] [CrossRef] [PubMed]
  106. Pellegrini, K.L.; Gerlach, C.V.; Craciun, F.L.; Ramachandran, K.; Bijol, V.; Kissick, H.T.; Vaidya, V.S. Application of small RNA sequencing to identify microRNAs in acute kidney injury and fibrosis. Toxicol. Appl. Pharmacol. 2016, 312, 42–52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Wu, Z.; Li, C.; Li, Q.; Li, J.; Lu, X. Puerarin alleviates cisplatin-induced acute renal damage and upregulates microRNA-31-related signaling. Exp. Ther. Med. 2020, 20, 3122–3129. [Google Scholar] [CrossRef]
  108. Joo, M.S.; Lee, C.G.; Koo, J.H.; Kim, S.G. miR-125b transcriptionally increased by Nrf2 inhibits AhR repressor, which protects kidney from cisplatin-induced injury. Cell Death Dis. 2013, 4, e899. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Zhang, H.; Zhang, X.; Yuan, X.; Wang, L.; Xiao, Y. MicroRNA-205 inhibits renal cells apoptosis via targeting CMTM4. Iran. J. Basic Med. Sci. 2015, 18, 1020–1026. [Google Scholar] [PubMed]
  110. Quintanilha, J.; Cursino, M.; Borges, J.; Torso, N.; Bastos, L.; Oliveira, J.; Cobaxo, T.; Pincinato, E.; Hirata, M.; Geraldo, M.; et al. MiR-3168, miR-6125, and miR-4718 as Potential Predictors of Cisplatin-Induced Nephrotoxicity in Patients with Head and Neck Cancer. Res. Sq. 2021. [Google Scholar] [CrossRef]
  111. Pavkovic, M.; Riefke, B.; Ellinger-Ziegelbauer, H. Urinary microRNA profiling for identification of biomarkers after cisplatin-induced kidney injury. Toxicology 2014, 324, 147–157. [Google Scholar] [CrossRef]
  112. Wolenski, F.S.; Shah, P.; Sano, T.; Shinozawa, T.; Bernard, H.; Gallacher, M.J.; Wyllie, S.D.; Varrone, G.; Cicia, L.A.; Carsillo, M.E.; et al. Identification of microRNA biomarker candidates in urine and plasma from rats with kidney or liver damage. J. Appl. Toxicol. 2017, 37, 278–286. [Google Scholar] [CrossRef] [PubMed]
  113. Glineur, S.F.; Hanon, E.; Dremier, S.; Snelling, S.; Berteau, C.; De Ron, P.; Nogueira da Costa, A. Assessment of a Urinary Kidney MicroRNA Panel as Potential Nephron Segment-Specific Biomarkers of Subacute Renal Toxicity in Preclinical Rat Models. Toxicol. Sci. 2018, 166, 409–419. [Google Scholar] [CrossRef]
  114. Kanki, M.; Moriguchi, A.; Sasaki, D.; Mitori, H.; Yamada, A.; Unami, A.; Miyamae, Y. Identification of urinary miRNA biomarkers for detecting cisplatin-induced proximal tubular injury in rats. Toxicology 2014, 324, 158–168. [Google Scholar] [CrossRef]
  115. Kagawa, T.; Zarybnicky, T.; Omi, T.; Shirai, Y.; Toyokuni, S.; Oda, S.; Yokoi, T. A scrutiny of circulating microRNA biomarkers for drug-induced tubular and glomerular injury in rats. Toxicology 2019, 415, 26–36. [Google Scholar] [CrossRef]
  116. Zhang, W.; Chen, C.; Jing, R.; Liu, T.; Liu, B. Remote Ischemic Preconditioning Protects Cisplatin-Induced Acute Kidney Injury through the PTEN/AKT Signaling Pathway. Oxid. Med. Cell Longev. 2019, 2019, 7629396. [Google Scholar] [CrossRef] [Green Version]
  117. Cho, Y.E.; Kim, S.H.; Lee, B.H.; Baek, M.C. Circulating Plasma and Exosomal microRNAs as Indicators of Drug-Induced Organ Injury in Rodent Models. Biomol. Ther. 2017, 25, 367–373. [Google Scholar] [CrossRef] [Green Version]
  118. Huang, S.J.; Huang, J.; Yan, Y.B.; Qiu, J.; Tan, R.Q.; Liu, Y.; Tian, Q.; Guan, L.; Niu, S.S.; Zhang, Y.; et al. The renoprotective effect of curcumin against cisplatin-induced acute kidney injury in mice: Involvement of miR-181a/PTEN axis. Ren. Fail. 2020, 42, 350–357. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Demethylation of IRF8 gene promoter during cisplatin-induced nephrotoxicity, contributing to renal tubular cell apoptosis. Created with BioRender.com accessed on 21 May 2021.
Figure 1. Demethylation of IRF8 gene promoter during cisplatin-induced nephrotoxicity, contributing to renal tubular cell apoptosis. Created with BioRender.com accessed on 21 May 2021.
Pharmaceuticals 14 00491 g001
Figure 2. Proposed role of histone modifications in the regulation involved in cisplatin-induced nephrotoxicity. Created with BioRender.com accessed on 21 May 2021.
Figure 2. Proposed role of histone modifications in the regulation involved in cisplatin-induced nephrotoxicity. Created with BioRender.com accessed on 21 May 2021.
Pharmaceuticals 14 00491 g002
Figure 3. Proposed model of miRs function in the activation of different signaling pathways during cisplatin-induced nephrotoxicity in rodent and human tubular cells. Block arrows indicate inhibitory effect; single arrows indicate stimulatory effect. The blue color indicates gene expression down-regulation, while the red color indicates gene expression up-regulation. Created with BioRender.com accessed on 21 May 2021.
Figure 3. Proposed model of miRs function in the activation of different signaling pathways during cisplatin-induced nephrotoxicity in rodent and human tubular cells. Block arrows indicate inhibitory effect; single arrows indicate stimulatory effect. The blue color indicates gene expression down-regulation, while the red color indicates gene expression up-regulation. Created with BioRender.com accessed on 21 May 2021.
Pharmaceuticals 14 00491 g003
Table 1. In vitro and in vivo evidence of microRNAs involved in cisplatin-induced nephrotoxicity in humans.
Table 1. In vitro and in vivo evidence of microRNAs involved in cisplatin-induced nephrotoxicity in humans.
miRNAAssayModel/Cell LineSourceCDDP DoseTimeEffectExp.Induced byGene TargetRef.
miR-9-3pin vitroHK-2 cell linePTECs2, 10 or 50 µM24 hToxicDown-CASK[84]
miR-18a-5pin vitroHPTEC cell linePTECs40 µM6 and 24 hToxicDown--[106]
miR-21in vivoPatients with malignant mesotheliomaUrine and kidneyCDDP chemotherapy4–144 hToxicUp-P21, BCL2[82]
in vitroHPTEC cell linePTECs85 µM24 hDown-P21, BCL2[82]
in vitrociPTEC-OAT1Cell culture media5 to 30 µM48 hToxicDown--[39]
miR-26ain vitroHK-2 cell linePTECs4 µmol/L24 hToxicDown-TRPC6, DRP1[25]
miR-29ain vitrociPTEC-OAT1Cell culture media5–30 µM48 hToxicUp--[39]
miR-30cin vitroHK-2 cell linePTECs10 µM24–72 hToxicDown-BNIP3L, HSPA5[22]
miR-31-5pin vitroHK-2 cell linePTECs40 µM0–48 hToxicUp-NUMB[107]
miR-34ain vitroHK-2 cell linePTECs9 µg/mL24 hToxicUp-SIRT1[37]
in vitrociPTEC-OAT1Cell culture media5–30 µM48 hToxicUp--[39]
miR-107in vitroHK-2 cell linePTECs7.5 mM6 hToxicUp-RPS19[83]
miR-125bin vitroHepG2, HEK-293 cell linePTECs30 µM24 hProtectiveUpNRF2AhRR, MDM2[108]
miR-140-5pin vitroHK-2 and 293T cell linesPTECs1.5–12 µM24 hToxicUp-NRF2, MnSOD, LDH[93]
miR-146b-5pin vitroHPTEC cell linePTECs40 µM6 and 24 hToxicDown--[106]
miR-181ain vitroHK-2 cell linePTECs50 µmol/L24 hToxicUp-BCL2, BAX[33]
miR-182in vitroHK-2 and 293T cell linePTECs20 µmol/L24 hToxicDown-FOXO1[15]
miR-192in vitrociPTEC-OAT1Cell culture media5–30 µM48 hToxicUp--[39]
miR-194-5pin vitroHK-2 cell linePTECs20 µM24 hToxicUp--[99]
miR-199a-3pin vitroHK-2 cell linePTECs100 µM24 hToxicUpp53mTOR[38]
miR-200cin vivo Patients with malignant mesotheliomaUrine and kidneyCDDP chemotherapy4–144 hToxicUp-P21, BCL2[82]
in vitroHPTEC cell linePTECs85 µM24 hDown-P21, BCL2[82]
miR-203in vitroHK-2 and 293T cell linePTECs20 µmol/L24 hToxicDown-GSK-3β[15]
miR-205in vitroHK-2 and HEK-293 cellsPTECs100 µg/mL6 hToxicDown-CMTM4[109]
miR-371b-5pin vitroHK-2 cell linePTECs2, 10 or 50 µM24 hToxicDown-CDK6[84]
miR-423in vivoPatients with malignant mesothelioma Urine and kidneyCDDP chemotherapy4–144 hToxicUp-P21, BCL2[82]
in vitroHPTEC cell linePTECs85 µM24 hDown-P21, BCL2[82]
miR-494in vitroHK-2 and 293T cell linePTECs20 µmol/L24 hToxicDown-ATF3[15]
miR-500a-3pin vitroHK-2 cell linePTECs20 µM24 hToxicDown-MLKL, RIPK1, RIPK3[99]
in vitroHK-2 cell linePTECs15 µM 6–48 hToxicDown-MLKL, RIPK3[98]
miR-577in vitroHK-2 cell linePTECs20 µM24 hToxicUp--[99]
miR-3168in vivoPatients with primary squamous cell carcinoma of the head and neckBloodCDDP chemotherapy-ToxicUp-PRKAB2, TOP2A[110]
miR-4718in vivoPatients with primary squamous cell carcinoma of the head and neckBloodCDDP chemotherapy-ToxicDown-PRKAB2, ERCC1[110]
miR-6125in vivoPatients with primary squamous cell carcinoma of the head and neckBloodCDDP chemotherapy-ToxicUp-ERCC1[110]
PTECs: Proximal Tubular Epithelial Cells; OAT1: Organic Anion Transporter 1.
Table 2. In vitro and in vivo evidence of microRNAs involved in cisplatin-induced nephrotoxicity in rodents.
Table 2. In vitro and in vivo evidence of microRNAs involved in cisplatin-induced nephrotoxicity in rodents.
miRNASpecieAssayModel/Cell LineSourceCDDP DoseTimeEffectExp.Induced byGene TargetRef.
miR-9-3pRatin vitroNRK52E cell linePTECs2, 10 or 50 µM24 hToxicDown-CASK, CDK6, JNK[84]
miR-15Ratin vivoMale Han Wistar ratBlood, urine and kidney1 or 3 mg/kg0–26 dToxicUp--[111]
miR-16Ratin vivoMale Han Wistar ratBlood, urine and kidney1 or 3 mg/kg0–26 dToxicUp-BTG2, PHLD3A[111]
miR-20aRatin vivoMale Han Wistar ratBlood, urine and kidney1 or 3 mg/kg0–26 dToxicUp-p21[111]
miR-21Mousein vivoAdult C57BL/6 miceUrine, blood and kidney15 mg/kg4 dToxicUp--[92]
in vitroMM55.K tubular cellsPTECs8 µg/mL48 h
miR-22Ratin vivoMale Sprague-Dawley ratsUrine2 or 5 mg/kg72 hToxicUp--[112]
miR-26aMousein vivoMale C57BL/6 miceKidney20 mg/kg3 dToxicDown-TRPC6, DRP1[25]
miR-26bRatin vivoMale Sprague-Dawley ratsUrine2 or 5 mg/kg72 hToxicUp--[112]
Male Wistar ratsBlood and kidney5 mg/kg3 dToxicDown-TGFβR-1, TAK1, mTOR, LC3-II[90]
miR-30cRatin vivoWistar ratKidney10 mg/kg1, 3 or 7 dToxicDown-BNIP3L, HSPA5[22]
in vitroNRK-52E cell linePTECs10 µM24–72 h
miR-31-5pRatin vivoFemale Sprague-Dawley ratsBlood and kidney20 mg/kg14 dToxicUp-NUMB[107]
miR-34aMousein vitroBUMPT-306 cell linePTECs40 µmol/L0–20 hToxicUp-p53[21]
Mousein vivoWild-type and p53-deficientC57BL/6 miceBlood and kidney30 mg/kg0–3 d
Mousein vivoMale C57BL/6 miceKidney15 mg/kg3 dToxicUp-SIRT1, p53, PHLDA3[36]
Ratin vitroNRK-52E cell linePTECs9 µg/mL24 hToxicUp-SIRT1[37]
RatMousein vivoMale Wistar ratsMale C57BL-6J miceBlood and kidney7 mg/kg25 mg/kg12 d--[37]
Ratin vivoMale Sprague-Dawley ratsKidney and plasma2 or 5 mg/kg72 hToxicUp--[112]
Ratin vivoMale Wistar ratsBlood and kidney5 mg/kg3 dToxicDown-TGFβR-1, TAK1, mTOR, LC3-II[90]
miR-34cRatin vivoMale Sprague-Dawley ratsKidney and plasma2 or 5 mg/kg72 hToxicUp--[112]
miR-34c-5pRatin vivoMale Hannover Wistar ratsKidney and urine2.5 mg/kg0–27 dToxicDown--[113]
miR-92aMousein vivoAdult C57BL/6 miceUrine, blood and kidney15 mg/kg4 dToxicDown--[92]
in vitroMM55.K tubular cellsPTECs8 µg/mL48 h-
miR-92bRatin vivoMale Sprague-Dawley ratsPlasma2 or 5 mg/kg72 hToxicUp--[112]
miR-93-5pRatin vivoMale Sprague-Dawley ratsKidney and urine1, 3 or 6 mg/kg1–7 dToxicUp--[114]
miR-107Ratin vivoMale Sprague-Dawley ratsKidney and blood6 mg/kg3 dToxicUp-RPS19[83]
miR-122Mousein vivoMale C57BL/6 miceKidney15 mg/kg3 dToxicDown-FOXO3, p53, PHLDA3[36]
miR-122-5pRatin vivoMale Sprague-Dawley ratsBlood and kidney6 mg/kg1–5 dToxicDown--[115]
miR-125bMousein vivoMale C57BL/6 mice and Nrf2 KOBlood and kidney15 mg/kg3 dProtectiveUpNRF2AhRR, MDM2[108]
miR-128Ratin vivoMale Sprague-Dawley ratsPlasma2 or 5 mg/kg72 hToxicUp--[112]
miR-130bRatin vivoMale Sprague-Dawley ratsPlasma2 or 5 mg/kg72 hToxicUp--[112]
miR-134Ratin vivoMale Sprague-Dawley ratsPlasma2 or 5 mg/kg72 hToxicUp--[112]
miR-138Mousein vivoFemale Diversity Outbred miceUrine and kidney5 mg/kg3 dToxicUp--[26]
miR-140Ratin vivoMale Sprague-Dawley ratsUrine2 or 5 mg/kg72 hToxicUp--[112]
miR-140-5pMousein vivoAdult male miceBlood and kidney20 mg/kg1–14 dToxicUp-NRF2, MnSOD, LDH[93]
in vivoMale C57BL/6J miceBlood and kidney20 mg/kg19 dToxicDown-NRF2, p53[24]
miR-141Mousein vivoAdult C57BL/6 miceUrine, blood and kidney15 mg/kg4 dToxicDown-ULK2[92]
in vitroMM55.K tubular cellsPTECs8 µg/mL48 h-
miR-143-3pRatin vivoMale Sprague-Dawley ratsBlood and kidney6 mg/kg1–5 dToxicDown--[115]
miR-144Mousein vivoMale C57BL/6 miceBlood and kidney20 mg/kg6 hToxicDown-PTEN, ATK, GSk3β[116]
Ratin vivoMale Sprague-Dawley ratsKidney2 or 5 mg/kg72 hToxicDown--[112]
miR-146aMousein vivoMale Balb/C miceBlood and kidney10 mg/kg24 hToxicUp--[117]
miR-146bRatin vivoSprague-Dawley ratsSerum and kidney6 mg/kg5 dToxicUp-ERBB4[89]
in vitroNRK-52E cell linePTECs7.5 µM48 h-
in vivoMale Sprague-Dawley ratsKidney2 or 5 mg/kg72 hToxicUp--[112]
miR-151Ratin vivoMale Sprague-Dawley ratsPlasma2 or 5 mg/kg72 hToxicUp--[112]
miR-155Mousein vivomiR-155−/− and C57BL/6 miceBlood and kidney20 mg/kg72 hToxicDown-c-FOS, TNF-α, IL-1β[100]
miR-181aMousein vivoMale C57BL/6 miceBlood and kidney20 mg/kg3 dToxicUp-PTEN[118]
miR-181bRatin vivoMale Sprague-Dawley ratsPlasma2 or 5 mg/kg72 hToxicUp--[112]
miR-182Mousein vivoMale C57BL/6 miceBlood and kidney30 mg/kg72 hToxicDown-FOXO1[15]
miR-191aRatin vivoMale Sprague-Dawley ratsPlasma2 or 5 mg/kg72 hToxicUp--[112]
miR-191a-5pRatin vivoMale Sprague-Dawley ratsKidney and urine1, 3 or 6 mg/kg1–7 dToxicUp--[114]
miR-192Ratin vivoMale Han Wistar ratBlood, urine and kidney1 or 3 mg/kg0–26 dToxicUp--[111]
miR-192-5pRatin vivoMale Sprague-Dawley ratsKidney and urine1, 3 or 6 mg/kg1–7 dToxicUp--[114]
Mousein vivoMale C57BL/6J miceKidney and blood20 mg/kg19 dToxicDown-NRF2, p53[24]
miR-193Ratin vivoMale Han Wistar ratBlood, urine and kidney1 or 3 mg/kg0–26 dToxicUp--[111]
miR-199a-3pMousein vivoMale C57/BL miceKidney20 mg/kg3 dToxicUpp53mTOR[38]
miR-203Mousein vivoMale C57BL/6 miceBlood and kidney30 mg/kg72 hToxicDown-GSK-3β[15]
miR-208aMousein vivoAdult C57BL/6 miceUrine, blood and kidney15 mg/kg4 dToxicDown--[92]
in vitroMM55.K tubular cellsPTECs8 µg/mL48 h-
miR-210Ratin vivoMale Han Wistar ratBlood, urine and kidney1 or 3 mg/kg0–26 dToxicUp--[111]
miR-218Mousein vivoFemale Diversity Outbred miceUrine and kidney5 mg/kg3 dToxicUp--[26]
miR-292-3pMousein vivoAdult C57BL/6 miceUrine, blood and kidney15 mg/kg4 dToxicDown--[92]
in vitroMM55.K tubular cellsPTECs8 µg/mL48 h-
miR-293*Mousein vivoAdult C57BL/6 miceUrine, blood and kidney15 mg/kg4 dToxicUp--[92]
in vitroMM55.K tubular cellsPTECs8 µg/mL48 h-
miR-371b-5pRatin vitroNRK52E cell linePTECs2, 10 or 50 µM24 hToxicDown-CASK, CDK6, JNK[84]
miR-375Mousein vivoWild-type and PT-Dicer−/− miceKidney30 mg/kg3 dToxicUpp53, NF-κBHNF-1β[85]
Ratin vitroRPTC cell linePTECs20 µM16 h
miR-377Mousein vivoAdult C57BL/6 miceUrine, blood and kidney15 mg/kg4 dToxicUp-CUL1[92]
in vitroMM55.K tubular cellsPTECs8 µg/mL48 h-
miR-378aRatin vivoMale Sprague-Dawley ratsUrine2 or 5 mg/kg72 hToxicUp--[112]
miR-431Ratin vivoMale Sprague-Dawley ratsPlasma2 or 5 mg/kg72 hToxicUp--[112]
miR-449Ratin vitroNRK-52E cell linePTECs20 µg/mL24 hToxicUp-SIRT1, P53, BAX[32]
miR-451Ratin vivoMale Sprague-Dawley ratsKidney2 or 5 mg/kg72 hToxicDown--[112]
miR-463Mousein vivoAdult C57BL/6 miceUrine, blood and kidney15 mg/kg4 dToxicUp--[92]
in vitroMM55.K tubular cellsPTECs8 µg/mL48 h-
miR-484Ratin vivoMale Sprague-Dawley ratsPlasma2 or 5 mg/kg72 hToxicUp--[112]
miR-494Mousein vivoMale C57BL/6 miceBlood and kidney30 mg/kg72 hToxicDown-ATF3, IL6[15]
miR-673-5pMousein vivoAdult C57BL/6 miceUrine, blood and kidney15 mg/kg4 dToxicDown--[92]
in vitroMM55.K tubular cellsPTECs8 µg/mL48 h-
miR-675-3pMousein vivoAdult C57BL/6 miceUrine, blood and kidney15 mg/kg4 dToxicUp--[92]
in vitroMM55.K tubular cellsPTECs8 µg/mL48 h-
miR-685Mousein vivoFemale Diversity Outbred miceUrine and kidney5 mg/kg3 dToxicUp--[26]
miR-702Ratin vivoMale Sprague-Dawley ratsPlasma2 or 5 mg/kg72 hToxicUp--[112]
miR-709Mousein vivoC57BL/6 miceBlood and kidney20 mg/kg72 hToxicUp-TFAM[23]
in vitromPTC cell linePTECs1–20 µM2–24 h-
miR-741Mousein vivoAdult C57BL/6 miceUrine, blood and kidney15 mg/kg4 dToxicUp--[92]
in vitroMM55.K tubular cellsPTECs8 µg/mL48 h-
miR-1190Mousein vivoAdult C57BL/6 miceUrine, blood and kidney15 mg/kg4 dToxicDown--[92]
in vitroMM55.K tubular cellsPTECs8 µg/mL48 h-
miR-1839Ratin vivoMale Sprague-Dawley ratsUrine2 or 5 mg/kg72 hToxicUp--[112]
miR-6215Ratin vivoMale Sprague-Dawley ratsPlasma2 or 5 mg/kg72 hToxicUp--[112]
miR-let-7bRatin vivoMale Wistar ratsBlood and kidney5 mg/kg3 dToxicDown-TGFβR-1, TAK1, mTOR, LC3-II[90]
miR-let-7eRatin vivoMale Sprague-Dawley ratsPlasma2 or 5 mg/kg72 hToxicUp--[112]
miR-let-7gRatin vivoMale Sprague-Dawley ratsUrine2 or 5 mg/kg72 hToxicUp--[112]
miR-let-7g-5pRatin vivoMale Sprague-Dawley ratsKidney and urine1, 3 or 6 mg/kg1–7 dToxicUp--[114]
PTECs: Proximal Tubular Epithelial Cells.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Loren, P.; Saavedra, N.; Saavedra, K.; Zambrano, T.; Moriel, P.; Salazar, L.A. Epigenetic Mechanisms Involved in Cisplatin-Induced Nephrotoxicity: An Update. Pharmaceuticals 2021, 14, 491. https://0-doi-org.brum.beds.ac.uk/10.3390/ph14060491

AMA Style

Loren P, Saavedra N, Saavedra K, Zambrano T, Moriel P, Salazar LA. Epigenetic Mechanisms Involved in Cisplatin-Induced Nephrotoxicity: An Update. Pharmaceuticals. 2021; 14(6):491. https://0-doi-org.brum.beds.ac.uk/10.3390/ph14060491

Chicago/Turabian Style

Loren, Pía, Nicolás Saavedra, Kathleen Saavedra, Tomás Zambrano, Patricia Moriel, and Luis A. Salazar. 2021. "Epigenetic Mechanisms Involved in Cisplatin-Induced Nephrotoxicity: An Update" Pharmaceuticals 14, no. 6: 491. https://0-doi-org.brum.beds.ac.uk/10.3390/ph14060491

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