Introduction

Cisplatin is a clinically advanced and highly effective anticancer drug that is used for the treatment of various solid tumors, such as lung cancer, stomach cancer, and ovarian cancer [1]. However, nephrotoxicity is the major side effect of cisplatin administration. Clinically, the risk of nephrotoxicity in patients taking cisplatin is between 20% and 35% and leads to death in acute kidney injury (AKI) patients [2, 3]. In addition, pediatric patients also develop nephrotoxicity when using cisplatin [4]. Patients with AKI are clinically characterized by impaired renal tubular function, acute renal failure, a reduction in whole blood cells, anemia, physical tremors, weight loss, gastrointestinal dysfunction, lethargy, and orbital tightening, which limit the antitumor use of cisplatin [5]. Cisplatin mediates nephrotoxicity via a number of different cytotoxic mechanisms. In addition to DNA damage, cisplatin also causes cytoplasmic organelle dysfunction, particularly in the endoplasmic reticulum and mitochondria, activates apoptotic pathways, and inflicts cellular damage via oxidative stress and inflammation [6].

Presently, there is no clinically effective drug to prevent or treat cisplatin-induced nephrotoxicity. Many high-efficacy and low-toxicity drugs from natural products have been developed to protect against cisplatin-induced AKI. For example, ginseng, curcumin, and pomegranate can act as antioxidants and anti-inflammatory agents and possibly protect against oxidative stress by restoring the levels of antioxidant enzymes [7]. In addition, pretreatment with vitamin supplements, such as vitamin E and riboflavin (vitamin B), significantly reduces serum urea and increases the expression levels of antioxidant enzymes in children with steroid-responsive nephrotic syndrome [8]. These natural products have potential antioxidant and anti-inflammatory properties and can be used as supplements to alleviate cisplatin-induced nephrotoxicity.

In this review, we first introduce the pathological manifestations of cisplatin-induced nephrotoxicity and clarify the molecular events of the underlying mechanisms. Finally, we summarize the roles of various kinds of natural products in protecting against cisplatin-induced AKI. This review focuses on the different mechanisms and protective effects of natural products, providing a comprehensive understanding of the prevention of cisplatin-induced nephrotoxicity and potential implications for drug combinations or natural supplements for AKI patients.

Pathological manifestations of cisplatin-induced nephrotoxicity

Clinically, different doses of cisplatin may lead to different degrees of nephrotoxicity. Patients who receive a single dose of cisplatin may suffer from reversible kidney injury, while large doses or multiple courses of treatment may cause irreversible renal failure [9]. Pharmacokinetic studies also show that nephrotoxicity is mainly due to the high volume of cisplatin distribution and long-term accumulation of cisplatin in the kidney [10]. In general, the pathological mechanisms of cisplatin-induced nephrotoxicity mainly manifest as decreases in renal blood flow and glomerular filtration rate [11] and ischemia or necrosis of proximal renal tubular epithelial cells [12].

Histopathological changes in cisplatin-induced nephrotoxicity are positively correlated with the dose of cisplatin. First, cisplatin is passively absorbed into renal tubular cells via organic cation transporter 2 (OCT2) and forms hydrates with water molecules, leading to continuous accumulation in renal cells [13]. The formation of cisplatin hydrate is a reversible process, and cisplatin hydrate can be dissociated into cisplatin and water molecules and discharged from the cells [13]. Thus, the accumulation and retention of cisplatin in renal cells leads to DNA damage, oxidative stress, apoptosis, and autophagy (Fig. 1).

Fig. 1: Schematic illustration of pathological manifestations of cisplatin-induced nephrotoxicity.
figure 1

The normal epithelium is damaged by cisplatin, as characterized by the loss of brush borders, epithelial cell necrosis, sloughing and obstruction, and immune cell infiltration.

Cisplatin first causes shedding of the brush shape of renal tubular epithelial cells. With increasing cisplatin accumulation, epithelial cells undergo necrosis and are gradually shed, accompanied by the formation of proteinaceous casts [14]. Moreover, the proximal tubule basement membrane becomes thickened, and tubules become dilated [15]. Electron microscopy observation of epithelial cell ultrastructure shows swollen and vacuolated mitochondria, endoplasmic reticulum expansion, and increased numbers of lysosomes [16]. Taken together, these organelle malfunctions result in the destruction and sloughing of epithelial cells, as well as the formation of intratubular obstructions.

Damaged renal tubular epithelial cells recruit many immune cells, such as macrophages, dendritic cells, and T cells, which release a variety of inflammatory factors [17]. Moreover, cisplatin can cause reduced medullary blood flow and exacerbate tubular cell injury, leading to acute ischemic injury in the kidneys [18]. Instead of the typical self-regulatory renal vasodilation in ischemic kidneys, evident vasoconstriction occurs in cisplatin-induced AKI, leading to hypoxic injury and vascular injury in severe cases [19]. Some studies have shown that cisplatin forms a complex with reduced glutathione in the liver and then enters the kidney. Cisplatin is decomposed into a nephrotoxic metabolite due to the action of glutamyltransferase in the brush edge of the renal proximal tubule, causing renal cell apoptosis or necrosis [20].

Mechanisms of cisplatin-induced nephrotoxicity

The application of cisplatin chemotherapy is often limited by severe adverse effects, including nephrotoxicity, ototoxicity, neurotoxicity, and vomiting. Nephrotoxicity, which is the major limiting factor of cisplatin use, involves various mechanisms, such as oxidative stress, apoptosis, inflammation, and autophagy (Fig. 2). Understanding the underlying mechanism is important for investigating intervention strategies for nephrotoxicity.

Fig. 2: The mechanism summary of cisplatin-induced nephrotoxicity.
figure 2

The mechanisms mainly include the transport and metabolism of cisplatin, apoptosis, autophagy, DNA damage, oxidative stress, and inflammation, which work together to aggravate AKI induced by cisplatin.

Cellular uptake and transport

Cisplatin is mainly excreted through the kidneys. It becomes concentrated during excretion, and the concentration in renal tubular epithelial cells is much higher than that in the blood. In the kidney, cisplatin is absorbed by renal cells via passive diffusion. During excretion, cisplatin and its metabolites are secreted and reabsorbed in the renal tubules during glomerular filtration, leading to a high concentration of cisplatin in the kidneys.

Recent studies have shown that cisplatin is taken up by renal tubular cells via OCT2, copper ion transporter 1 (CTR1), and solute carrier family 22 member 2 [21]. In addition, cisplatin is secreted into the lumen by solute carrier family 47 member 1 and multidrug and toxin extrusion 1 [22]. Knockdown of the Oct2 gene can significantly reduce cisplatin-induced nephrotoxicity [23]. Consistently, patients with Oct2 mutations show low OCT2 expression and reduced cisplatin transport into renal tubular cells, resulting in decreased nephrotoxicity [24]. In addition, when CTR1 expression is downregulated, cisplatin uptake and the subsequent cytotoxicity decrease significantly [25]. Moreover, peroxiredoxin I (Prx I)-deficient mice have higher resistance to cisplatin-induced nephrotoxicity than wild-type mice due to increased cisplatin excretion via the high expression of the renal efflux transporters multidrug resistance-related protein 2 (MRP2) and MRP4 in Prx I-deficient mice [26].

DNA damage

Cisplatin mediates its cytotoxic effects by binding DNA to form adducts that cause DNA damage [27]. In an aqueous environment, the chloride ligand of cisplatin is replaced by water molecules to form a positively charged hydrated complex ion, which is transferred to the nucleus by DNA electrostatic attraction. Then, this complex binds to DNA to form an adduct, resulting in DNA cross-linking and preventing DNA synthesis and replication in rapidly proliferating cells [28]. This phenomenon is pronounced in cells with defective DNA repair.

However, cisplatin binds nonspecifically to nuclear DNA, and less than 1% of platinum binds to nuclear DNA [29]. Interestingly, mitochondrial DNA is more sensitive than nuclear DNA to cisplatin-mediated cytotoxicity [30]. The positively charged metabolites produced by the hydrolysis of cisplatin preferentially accumulate in mitochondria, which are negatively charged. Therefore, the sensitivity of cells to cisplatin depends on mitochondrial density and the mitochondrial membrane potential in cells [31]. Given that the renal proximal tubule contains sites of quite high mitochondrial density, it is the most highly sensitive site in the kidney to cisplatin [32].

Apoptosis

It has been reported that a low concentration (8 μM) of cisplatin causes renal tubular epithelial apoptosis, while a high concentration (800 μM) of cisplatin induces necrosis [33]. Cisplatin-induced apoptosis in renal tubular cells is primarily associated with mitochondria-mediated endogenous pathways, death receptor-mediated exogenous pathways, and endoplasmic reticulum stress (ERS) pathways.

Mitochondria-mediated endogenous pathways

Cisplatin-induced mitochondria-mediated apoptotic pathways mainly include caspase-dependent and -independent pathways. When cisplatin enters renal tubular epithelial cells, BAX translocates to mitochondria and activates caspase-2, resulting in the release of cytochrome c, second mitochondria-derived activator of caspase/direct inhibitor of apoptosis proteins binding protein with low Pi (isoelectric point) (SMAC/DIABLO), high temperature requirement A2 (HtrA2/Omi), and apoptosis-inducing factor (AIF) from mitochondria [34]. Then, caspase-9 is activated, which eventually leads to apoptosis [35]. Apart from the caspase-dependent pathway, cytoplasmic Omi/HtrA2 also promotes caspase-independent apoptosis by binding and cleaving inhibitors of apoptotic proteins after cisplatin-induced apoptotic stimulation [36].

AIF is an apoptosis-related protein located on the mitochondrial membrane, and poly (ADP-ribose) polymerase-1 (PARP-1) is a nuclear factor that participates in DNA repair and protein modification. Once cellular DNA is severely damaged by cisplatin, nuclear PARP-1 activity is increased, causing AIF activation and nuclear translocation, which induces apoptosis [37]. PARP-1 activation is a primary signal in the process of cisplatin-induced nephrotoxicity. Moreover, PARP-1 inhibition or deletion protects the kidneys from nephrotoxicity, providing a therapeutic strategy for cisplatin-induced nephrotoxicity [38].

The role of p53 in cisplatin-induced cytotoxicity mainly involves activation of the mitochondrial pathway. Upon exposure to cisplatin-induced cellular DNA damage, p53 is phosphorylated, and the proapoptotic protein BAX undergoes structural modifications and alters mitochondrial membrane integrity, causing the activation of p53 upregulated modulator of apoptosis-α and Ca2+-independent phospholipase A2. Then, the antiapoptotic proteins BCL-2 and BCL-XL are downregulated, triggering the mitochondrial apoptotic pathway [39].

Death receptor-mediated exogenous pathways

In the exogenous apoptotic pathways, cisplatin binds to death receptors such as tumor necrosis factor receptor 1 (TNFR1), TNFR2, and FAS on the cell membrane to activate caspase-8, which further activates caspase-3, ultimately leading to apoptosis [40]. Cisplatin upregulates the expression of tumor necrosis factor-α (TNF-α), promoting the interaction of TNF-α and TNF receptors, including TNFR1 and TNFR2. TNFR1 has a death domain and is able to directly trigger exogenous apoptosis. However, TNFR2 mainly regulates the inflammatory response to induce apoptosis because it has no death domain [41]. In addition, cisplatin can also activate the FAS/FAS-L system [42], and the FAS-associated death domain further interacts with FAS or TNFR1 to trigger apoptosis, but the detailed mechanisms have not been elucidated.

Endoplasmic reticulum stress pathways

Cisplatin can also activate the apoptotic pathway that is mediated by ERS. After cisplatin enters cells, it acts on the cytochrome P450 (CYP450) enzymatic system on the endoplasmic reticulum membrane to induce oxidative stress and activate caspase-12, which leads to apoptosis [43]. As expected, cisplatin-induced apoptosis is significantly reduced in cytochrome P450, family 2, subfamily E, polypeptide 1 (Cyp2e1)-knockout mice [44]. Similarly, another study showed that the expression of the ERS marker X-box-binding protein 1 was increased, and calpain and caspase-12 cleavage products were observed in rat kidneys after cisplatin treatment [45]. Furthermore, transfection with an anti-caspase-12 antibody significantly attenuated cisplatin-induced apoptosis in porcine kidney LLC-PK1 cells [46]. The ERS pathway is also involved in the activation of endoplasmic reticulum phospholipase A2, which limits downstream p53 and activates upstream caspase-3. The endoplasmic reticulum may be a link between p53 and caspase-3 in the absence of mitochondrial dysfunction [47].

Oxidative stress

In recent years, studies have shown that oxidative stress and nitrosative stress play vital roles in cisplatin-induced nephrotoxicity, which is characterized by increased malondialdehyde (MDA), 4-hydroxy, 8-hydroxydeoxyguanosine, and 3-nitrotyrosine, and decreased superoxide dismutase (SOD) and catalase (CAT) after cisplatin treatment. Thus, reactive oxygen species (ROS) scavengers and antioxidants show robust protective effects against nephrotoxicity [48].

After entering renal tubular cells, cisplatin can rapidly react with the thiol-containing antioxidants glutathione and metallothionein to degrade or inactivate them. Moreover, some antioxidant enzymes, such as glutathione peroxidase, SOD, and glutathione reductase, are also inhibited, leading to increased ROS levels [49]. ROS affect the activity of mitochondrial complex enzymes I–IV, thereby inhibiting the normal transmission of the oxidative respiratory chain and leading to adenosine triphosphate depletion [40]. Then, increased ROS results in lipid peroxidation, changing membrane structure and permeability, which further affect cellular function [50]. Finally, ROS impair amino acids, proteins, and carbohydrates, thus promoting DNA damage and apoptosis. In addition, increased ROS can induce increased expression of FAS-L, FAS, TNFR1, and TNF-α, eventually resulting in apoptosis [45].

Inflammation

Cisplatin-induced nephrotoxicity is associated with the inflammatory response. Renal TNF-α expression is increased in a cisplatin-induced nephrotoxic mouse, and cisplatin-induced renal insufficiency and injury can be significantly alleviated by TNF-α inhibition or knockout, indicating that increased TNF-α expression plays an important role in cisplatin-induced nephrotoxicity [51]. Interestingly, after cisplatin administration, TNF-α in the circulation and urine may be derived from renal epithelial cells rather than immune cells. Moreover, TNF-α induces the production of ROS, further activating the transcription factor, nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB), which in turn induces the production of proinflammatory cytokines such as TNF-α [52]. The inhibition of NF-κB transcriptional activity by JSH-23 (a kind of NF-κB inhibitor) improves kidney function in mice [53].

TNF-α activates proinflammatory cytokines and chemokines to trigger oxidative stress, ultimately exacerbating kidney damage. Hydroxyl free radicals produced by cisplatin are involved in the phosphorylation of p38 mitogen-activated protein kinase (p38 MAPK) and the regulation of TNF-α synthesis, ultimately inducing the activation of NF-κB. Therefore, the hydroxyl radical scavenger dimethyl thiourea inhibits p38 MAPK activation and TNF-α mRNA expression in murine kidneys. The inhibition of p38 MAPK reduces the production of TNF-α, thereby effectively protecting against cisplatin-induced kidney damage [54]. Other cytokines, such as transforming growth factor-β, monocyte chemoattractant protein-1 (MCP-1), intercellular adhesion molecule, and heme oxygenase-1 (HO-1), are also associated with cisplatin-induced nephrotoxicity [55]. N-Acetylcysteine (NAC), an antioxidative agent, effectively inhibits inflammation and activation of the complement system to exert renal protection [56]. Mitochondrial dysfunction leads to the formation of O2–, while the inflammatory response induced by cisplatin involves the upregulation of TNF-α, nicotinamide adenine dinucleotide phosphate oxidase, and inducible nitric oxide synthase (iNOS), which directly leads to NO formation. NO and O2– produce ONOO, which has strong oxidation and nitration properties, further inducing apoptosis and necrosis [57].

Autophagy

Autophagy plays an important role in maintaining cellular homeostasis and surviving cisplatin-induced nephrotoxicity. In NRK-52E cells treated with cisplatin, the increases in autophagy and apoptosis were both inhibited after beclin-1 knockdown, indicating that autophagy mediates cell damage [58]. However, another study showed that autophagy inhibition accelerated apoptosis, demonstrating the protective effect of autophagy in cisplatin-induced kidney injury [59]. Moreover, autophagy can prevent AKI and proximal tubule apoptosis caused by cisplatin [60].

Studies have reported that the suppression of mammalian target of the rapamycin (mTOR) activity alleviates the inhibitory phosphorylation of Unc-51-like autophagy activating kinase 1, which leads to the activation of autophagy [61]. Pretreatment with rapamycin, an mTOR inhibitor, induces autophagy to improve renal function in rats with ischemia/reperfusion [62]. Interestingly, NAD(P)H quinone dehydrogenase 1 deletion (an oxidative stress barrier) enhances the effect of rapamycin and leads to increased tuberous sclerosis complex 2 phosphorylation, indicating that autophagy may be activated to counter the increased stress and protect against AKI [63].

Current treatment of cisplatin-induced nephrotoxicity

Various treatments have been applied to address the different mechanisms of cisplatin-induced nephrotoxicity (Table 1). For example, cimetidine acts as an OCT2 inhibitor that inhibits the transportation of cisplatin in the kidney to protect against AKI [64], carvedilol works as an antioxidant against the oxidative stress process [65], cilastatin inhibits the apoptotic pathway [66], and rosiglitazone reduces inflammation [67].

Table 1 Current treatments for cisplatin-induced nephrotoxicity.

At present, although several kinds of drugs are applied clinically in response to kidney damage caused by cisplatin, these drugs exhibit different degrees of inadequacy. For example, hydration and diuresis in the clinic enhance cisplatin excretion and reduce renal exposure [68]. However, the disadvantage is that a large amount of hydration is required before and after cisplatin administration [69]. Moreover, adverse reactions such as osmotic pressure changes may occur during chemoprevention. In addition, metabolic waste in the body can be excreted through hemodialysis, which is often accompanied by hypophosphatemia and heart rate disorders. Amifostine is a broad-spectrum cytoprotective agent approved by the FDA as a kidney protectant for cisplatin chemotherapy in patients with advanced ovarian cancer; however, its application in other tumors is limited due to blood pressure drops and hypocalcemia [70].

Protective effects of natural products that prevent cisplatin-induced nephrotoxicity

Traditional and complementary medicines, including a variety of natural products, such as herbs, vitamins, minerals, trace elements, and nutritional supplements, have been widely used in most countries [71]. Adopting natural products in healthcare can improve the physical fitness of patients. To better understand the roles of natural products in AKI, we summarized the protective effects of various classes of natural products on cisplatin-induced nephrotoxicity (Fig. 3 and Tables 2 and 3).

Fig. 3: The summary of natural products to protect against cisplatin-induced nephrotoxicity.
figure 3

Potential natural product treatments for cisplatin-induced nephrotoxicity classified by chemical structures.

Table 2 Natural products in the treatments for cisplatin-induced nephrotoxicity classified by chemical structures.
Table 3 Natural products in the treatments for cisplatin-induced nephrotoxicity classified by mechanisms.

Flavonoids

Studies have shown that formononetin can effectively reduce OCT2 expression and increase MRP expression, resulting in decreased accumulation of cisplatin in renal tubular cells [72]. Similarly, puerarin protects against cisplatin-induced nephrotoxicity and promotes the antitumor activity of cisplatin in COLO205 and HeLa tumor cells in a dose-dependent manner [73]. Interestingly, naringin can alleviate cisplatin-induced renal dysfunction by inhibiting the inflammatory response and reducing apoptosis [74]. Flavonoids with multiple activities, such as icariin, breviscapine, epicatechin and epicatechin gallate, sappanone A, morin and its hydrate, quercetin, silymarin, daidzein, and xanthohumol, can reduce cisplatin-induced oxidative and nitrosative stress and decrease creatinine (Cre) and blood urea nitrogen (BUN) levels to improve renal function, thereby alleviating cisplatin-induced nephrotoxicity [75,76,77,78,79,80,81,82,83,84]. In addition, wogonin markedly inhibits receptor-interacting protein kinase 1-mediated necrosis and the canonical WNT pathway (WNT/β-catenin pathway) to protect against cisplatin-induced nephrotoxicity [85]. Further studies demonstrated that baicalein and apigenin ameliorated cisplatin-induced renal damage through the upregulation of antioxidant pathways and downregulation of the MAPK and NF-κB signaling pathways [86].

Interestingly, Scutellaria baicalensis Georgi not only enhances the therapeutic efficacy of cisplatin but also attenuates chemotherapy-induced AKI [87]. Glycyrrhizic acid, 18β-glycyrrhetinic acid, hypericin, and eriodictyol reduce AKI by inhibiting the cisplatin-induced phosphorylation of NF-κB and upregulating the expression of nuclear factor erythroid 2 (NFE2)-related factor 2 (NRF2) and HO-1 [88,89,90]. D-Pinitol and mangiferin attenuate inflammatory infiltration, DNA damage, and renal dysfunction in rats by modulating the MAPK pathway [91]. Furthermore, cisplatin-induced oxidative stress is mitigated by hesperidin and hesperetin by reducing MDA/Myeloperoxidase (MPO) levels and increasing SOD/Glutathione (GSH) levels. Galangin and the isoflavonoid biochanin A exhibit renoprotective effects in mice by targeting the inflammatory response and p53-mediated apoptosis. Importantly, luteolin significantly reduces histological and biochemical changes induced by cisplatin by blocking platinum accumulation and inflammation [92]. Genistein and naringin inhibit the NF-κB and iNOS pathways and p53 activation to improve HK-2 cell viability and kidney morphology in the presence of cisplatin and have become a potential effective treatment strategy for AKI [93]. A recent study demonstrated that scutellarin and anthocyanin from the fruits of Panax ginseng attenuate cisplatin-induced nephrotoxicity by inhibiting TNF-α [94]. In summary, flavonoids exhibit great potential as dietary supplements to ameliorate cisplatin-induced nephrotoxicity.

It is worth noting that the flavonoid phloretin is a robust toxicant (LC50 = 362 μM) that potentiates H2O2-induced toxicity, which is consistent with the previously noted cytotoxicity of phloretin and other hydroxychalcones. This toxicity is due to the oxidative activities of these polyphenols and the possible induction of mitochondrial toxicity [95].

Many flavonoids show strong protective effects against cisplatin-induced AKI. To date, researchers have found that many kinds of flavonoids activate NRF2/HO-1 signaling and inhibit NF-κB activity to alleviate kidney injury. More interestingly, some flavonoids not only protect against cisplatin-induced kidney injury but also synergistically inhibit the growth of tumors, enhancing the efficacy of cisplatin in tumor-bearing mice [74].  These results suggest that flavonoids may be used in the comprehensive treatment of cancer patients. Although flavonoids exhibit strong protection against kidney injury, there are some challenges in the clinical application of flavonoids. For example, monomers of flavonoid compounds are difficult to extract and have poor lipid solubility and low bioavailability, limiting their clinical applications [96]. If researchers can overcome these challenges, flavonoids will become promising drugs for AKI treatment.

Saponins

Oxidative stress and inflammation are important mechanisms involved in the pathogenesis of AKI. Some studies have shown that saikosaponin D can increase the survival rate of HK-2 cells and maintain the normal morphology of the nucleus. Saikosaponin D can inhibit the activation of the NF-κB-P38-JNK-MAPK signaling cascade, thereby reducing cisplatin-induced apoptosis [97]. Red ginseng, ginsenoside Rg5, and Platycodon grandiflorum saponins can inhibit inflammation by reducing the expression of cyclooxygenase-2 and iNOS to inhibit acute tubular necrosis and apoptosis [98, 99]. Renal oxidative stress, as evidenced by increased MDA levels and declines in GSH and SOD activities, is significantly reduced by saponins from Terminalia arjuna [100].

In addition, some saponin components mainly regulate autophagy and apoptosis to exert protective effects against kidney injury. Ginsenoside 20(S)-Rg3 and ginsenoside Rb3 can inhibit autophagy to improve renal injury by blocking the JNK-P53-caspase-3 signaling cascade [101, 153,154,155,156]. Recent reports show that pretreatment with curcumin can ameliorate cisplatin-induced kidney damage by suppressing inflammation and apoptosis [157].

The mechanisms of cisplatin-induced kidney damage involve various pathways, such as inflammatory mediators, oxidative stress, necrosis and apoptosis, and autophagy. To date, researchers have not found that these mechanisms are involved in cisplatin-induced nephrotoxicity, starting with excess ROS generation, which leads to oxidative stress, triggering inflammatory and autophagy pathways that damage DNA and induce apoptosis in the kidney. It is still unclear how the various pathways integrate and ultimately lead to kidney damage. In recent years, many natural products have been discovered by different mechanisms. A natural compound may have multiple active targets rather than only one unique target. Therefore, a natural product may play multiple roles and exhibit wide use and may have increased potential toxicity or side effects. Since some pathways of cisplatin-induced kidney injury are also involved in the antitumor effects of cisplatin, natural products may also affect cisplatin-mediated antitumor effects. While most compounds have anti-inflammatory and antioxidant properties, NAC and vitamin E have been reported to act as antioxidants and contribute to the development of lung cancer [158]. Therefore, it is unclear whether natural compounds with antioxidant activity interfere with the development of tumors while protecting against kidney injury. In this case, in addition to check the protective role against AKI, it is necessary to further study if natural products have effects on tumor growth, which may help to break through the limited use of cisplatin in clinic.

It is worth noting that natural products that have robust therapeutic effects on cisplatin-induced AKI also alleviate kidney diseases caused by other factors. Further research is needed to verify the beneficial effects of certain products on humans and other animals with kidney diseases to elucidate the detailed mechanisms of the renoprotective effects. To achieve the desired protective effect against nephrotoxicity, researchers should take all aspects of the relevant mechanisms into account and consider comprehensive measures or combinations of drugs. In addition, although certain natural products are excellent in protecting against kidney damage in vitro and in vivo, it is necessary to study the optimal dose for protecting against different tumors and different cisplatin strengths.

Furthermore, the development of molecular biology technology has led to the research of targeted therapy using cisplatin and natural products or derivatives that are highly selective for the kidney or tumor as carriers, and chemically coupling these factors into biological treatments. Direct delivery of cisplatin to the tumor site rather than the kidney can not only reduce the amount of cisplatin needed but also improve the efficacy and reduce adverse reactions. This opens up new ideas for the study of protective measures against cisplatin-induced nephrotoxicity.