Next Article in Journal
Valoration of the Synthetic Antioxidant Tris-(Diterbutyl-Phenol)-Phosphite (Irgafos P-168) from Industrial Wastewater and Application in Polypropylene Matrices to Minimize Its Thermal Degradation
Next Article in Special Issue
Skin Pigmentation Types, Causes and Treatment—A Review
Previous Article in Journal
One-Pot Synthesis of Polynuclear Indole Derivatives by Friedel–Crafts Alkylation of γ-Hydroxybutyrolactams
Previous Article in Special Issue
Potential Therapeutic Value of the STING Inhibitors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Endoplasmic Reticulum Stress and Mitochondrial Stress in Drug-Induced Liver Injury

College of Veterinary Medicine, Gansu Agricultural University, Lanzhou 730070, China
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(7), 3160; https://doi.org/10.3390/molecules28073160
Submission received: 11 February 2023 / Revised: 26 March 2023 / Accepted: 29 March 2023 / Published: 2 April 2023

Abstract

:
Drug-induced liver injury (DILI) is a widespread and harmful disease closely linked to mitochondrial and endoplasmic reticulum stress (ERS). Globally, severe drug-induced hepatitis, cirrhosis, and liver cancer are the primary causes of liver-related morbidity and mortality. A hallmark of DILI is ERS and changes in mitochondrial morphology and function, which increase the production of reactive oxygen species (ROS) in a vicious cycle of mutually reinforcing stress responses. Several pathways are maladapted to maintain homeostasis during DILI. Here, we discuss the processes of liver injury caused by several types of drugs that induce hepatocyte stress, focusing primarily on DILI by ERS and mitochondrial stress. Importantly, both ERS and mitochondrial stress are mediated by the overproduction of ROS, destruction of Ca2+ homeostasis, and unfolded protein response (UPR). Additionally, we review new pathways and potential pharmacological targets for DILI to highlight new possibilities for DILI treatment and mitigation.

1. Introduction

Drug-induced liver injury (DILI) refers to liver toxicity induced by a drug or its metabolite and is commonly categorized into intrinsic and idiosyncratic DILI [1,2]. Non-steroidal anti-inflammatory drugs, anti-tuberculosis drugs, antiepileptic drugs, traditional Chinese medicine (TCM), and other drugs can cause varying degrees of DILI. Due to the complex pathogenic mechanism of DILI, there is no current specific drug. [3,4,5,6,7,8].
DILI is tightly linked to mitochondrial and endoplasmic reticulum stress (ERS). Most proteins are synthesized, processed, and transported in the endoplasmic reticulum, a determinant compartment for ensuring cell homeostasis. Upon stimulation by drugs and their metabolites, the endoplasmic reticulum homeostasis is disturbed, which initiates ERS and can activate apoptosis pathways. This cascade of events potentiates the development of liver injury, causing liver failure in severe cases [9]. Mitochondria provide direct energy for cell function, maintain the environmental balance in hepatocytes, and are the target organelles of oxidative stress injury. Mitochondrial stress, caused by unfolded or misfolded proteins in the mitochondria, alters the permeability of the mitochondrial membrane, which subsequently leads to excessive ATP consumption, the release of reactive oxygen species (ROS), and impaired calcium homeostasis. These events result in mitochondrial swelling and the consequent release of pre-apoptotic factors, leading to hepatocyte apoptosis [10,11].
Communication between the ERS and mitochondrial stress is also an important topic of investigation. When cells undergo mitochondrial stress, the redox balance in the endoplasmic reticulum is disrupted, interfering with the endoplasmic reticulum function and triggering ERS [12]. Xiao et al. showed that drug-induced ERS could stimulate the expression of the mitochondrial stress factor HSP60, impair mitochondrial respiration, and decrease mitochondrial membrane potential in mouse hepatocytes [13]. This proves that ERS can also induce mitochondrial stress.
DILI has a high prevalence worldwide; however, improvements in the therapeutic approaches for this event are still lacking. Recent studies have identified new potential targets for the treatment of DILI, such as sphingosine kinase-1 (SPHK1) [14], steroidogenic acute regulatory protein (STARD1) [15], transient receptor potential 2 (TRPM2) [16], and activating transcription factor 4 (ATF4) [17]; genes associated with ERS and mitochondrial stress. Notably, inhibition of the SPHK1–STARD pathway effectively alleviates DILI. In this review, we discussed the mechanisms of liver injury caused by several types of drugs, focusing primarily on DILI caused by ERS and mitochondrial stress. Importantly, we highlight the main cellular events that mediate ERS and mitochondrial stress, such as the overproduction of ROS, impaired Ca2+ homeostasis, and UPR (unfolded protein response). Additionally, we highlight new (and old) pharmacological targets, aiming to contribute to the development of DILI treatment and mitigation.

2. Classification of Drugs Producing DILI and Their Associated Mechanisms in the Endoplasmic Reticulum and Mitochondrial Stress

2.1. DILI Classification According to the Liver Lesion and Drug Type

Generally, DILI is divided into intrinsic and idiosyncratic types, but it can also be classified according to the lesion and drug. The pathological targets of DILI involve hepatocyte injury (67.3%), bile duct epithelium damage (23.9%), and vascular injury (8.8%) [18]. Hepatocyte injury includes mitochondrial damage, calcium imbalance, and excessive ROS production, the majority of which occur due to lobular hepatitis; bile duct epithelial damage, manifestations of cholestatic patterns, a hepatocellular injury, vascular injury, including the obliteration of portal venules, and veno-occlusive disease or sinusoidal obstruction syndrome may also occur. According to the developmental processes of DILI, mild to severe diseases involve mild liver enzyme elevation, jaundice, hepatic steatosis, hepatitis, liver fibrosis, cirrhosis, hepatic portal vein sclerosis, liver failure, angiosarcoma, liver cancer, and death [19]. Mild diseases, if not treated timely, may deteriorate to severe forms or cause death [20].
The drugs that can cause DILI can be divided into antibiotics, non-steroidal anti-inflammatory drugs, antipsychotics, antidepressants, lipid-lowering drugs, anti-tuberculosis drugs, antiepileptic drugs, traditional Chinese medicine (TCM), dietary supplements, oral anticoagulants, anti-androgen drugs, antibacterial drugs, hypoglycemic drugs, and acid-suppressing drugs. The extension and type of lesions differ with each drug (Table 1). Nonetheless, some mechanisms involved in DILI development with certain drugs remain unclear. Currently, the drugs known to induce ERS and mitochondrial stress include acetaminophen (APAP), isoniazid (INH), and Valproic acid (VPA) [5,8,21,22].

2.2. Endoplasmic Reticulum and Mitochondrial Stress in Drug-Induced Liver Injury

2.2.1. ERS and ERS-Related Molecules in the Development of DILI

The endoplasmic reticulum (ER) comprises a nuclear membrane domain, smooth ER, rough ER, and regions of connection with other organelles. Peptide chains are processed and transported through the ER to the Golgi apparatus [56]. After the reticular Golgi apparatus processes and sorts the peptide chains, the resulting protein with a native conformation is expelled into the extracellular space. One of the main functions of the ER is to store calcium, which is released upon various signaling events in a controlled manner. The subsequent increase in cytoplasmic Ca2+ results in signal transduction [57]. The ER encompasses a quality control system that deals with excessive protein accumulation, protein misfolding, or continuous loss of Ca2+, which ensures cell death upon failure to adapt to a stressor.
When misfolded proteins accumulate, cells can degrade them via ER-associated protein degradation (ERAD) [58]. The ERAD involves the cytoplasmic ATPase p97, through which a membrane reverse-transcription translocation complex provides a pathway for the protein to return to the cytoplasm. Dislocated ERAD substrates are ubiquitinated at the cytoplasmic side of the ER membrane and targeted for proteasomal degradation [9]. When the ER cannot deal with the high load of protein folding, Endoplasmic reticulum stress (ERS) ensues, triggering the UPR, which enhances the protein processing ability of the ER. The UPR helps refold misfolded proteins in the ER, allowing its return to homeostasis. If the stress signal is severe or maintained, ER initiates the apoptotic pathway to remove the misfolded proteins and restore homeostasis [59].
  • APAP-induced endoplasmic reticulum stress
After APAP enters the liver, 90% is converted into non-toxic compounds under the influence of UGTs (glucuronidase) and SULTs (sulfur transferase) and excreted in the urine. The remaining 10% is metabolized by the cytochrome P450 (CYP450) enzyme system to form the metabolite N-acetyl-p-benzoquinonimine (NAPQI). In the normal (therapeutic doses) condition, NAPQI is detoxified with glutathione (GSH) [60]. However, when APAP ingestion reaches a toxic dose, a large amount of APAP enters the CYP450 pathway, and NAPQI is produced in larger amounts [49] (Figure 1). Once GSH is depleted, NAPQI binds to cellular proteins, such as mitochondrial protein, which initiates hepatocellular necrosis and DNA fragmentation [61]. Additionally, when GSH content decreases, the ER lumen produces a redox shift, protein disulfide isomerase (PDI) undergoes oxidation, and the redox imbalance damages the function of ER oxidoreductase, which may initiate the ER-related signaling pathway [62,63], This event activates endoplasmic reticulum-dependent signal transduction and apoptosis, which may be a leading mechanism of ERS [64]. Recent reports suggest that ERS and ERS-related molecules participate in developing APAP-induced liver injury. Nagy et al. showed significant induction of an ERS-responsive proapoptotic transcription factor, GADD153/CHOP (growth arrest and DNA damage-inducible gene 153 or C/EBP-homologous Protein), in a mouse model of APAP-induced liver injury [65]. CHOP participates in APAP-induced ERS during liver injury [66]. A recent report also suggested that upregulation of a mitochondrial cholesterol transporter, steroidogenic acute regulatory protein (STARD1), following ERS mediates APAP-induced hepatoxicity via mitochondrial SH3BP5 (SAB) and JNK1 (c-jun N-terminal kinase 1) and 2 phosphorylation [67]. Li et al. found that SPHK1 levels in the liver significantly increased after APAP treatment. Phosphorylation of SPHK1 can catalyze the formation of S1P (Sphingosine-1-phosphate) (Figure 2), which can activate PERK-eIF2α-ATF4 and ATF6 (Activating Transcription Factor 6) and induce the generation of the apoptosis signal CHOP, prompting ERS [14]. These experimental findings suggest that ERS and its related molecular events occur in liver injury via the regulation of chemical chaperons. Wang et al. demonstrated that INH-induced liver injury shares the same mechanisms as that of APAP, primarily by ONOO generated by oxidative stress.

2.2.2. Mitochondrial Stress and Mitochondrial Stress-Related Molecules Participate in the Development of DILI

Mitochondria consist of a mitochondrial inner membrane (IMM), mitochondrial membrane gap, mitochondrial outer membrane (OMM), and the mitochondrial matrix. In cells, mitochondria are the main source of NADH and are responsible for parts of the pyrimidine and lipid biosynthesis pathways, including the fatty acid β-oxidation pathway [68]. Similar to the ER, unfolded or misfolded proteins in the mitochondria increase mitochondrial pressure, resulting in mitochondrial stress. It was shown that loss of mitochondrial membrane potential, decreased canonical fusion proteins, and alterations in mitochondrial lipid composition can drive changes in mitochondrial morphology and that these changes lead to mitochondrial fragmentation and reduced function [69]. Mitochondrial dysfunction is associated with several human diseases, such as metabolic syndrome, cancer, and neurodegenerative diseases. Mitochondrial dysfunction also plays a key role in the pathogenesis of DILI, a consequence of altered metabolic pathways and damaged mitochondrial components [10,11]. When the mitochondrial death critical threshold is surpassed, mitochondrial damage can trigger liver necrosis or failure leading to the activation of cell death signaling pathways [10,70]. More specifically, sodium valproate impairs mitochondrial respiration, aspirin affects mitochondrial β-oxidation, and diclofenac damages the mitochondrial membrane.
  • APAP-induced mitochondrial stress
At higher APAP doses, mitochondrial shape has a biphasic response. The early changes in mitochondrial morphology are reversible and help pre-serve mitochondrial function. In contrast, late delayed change in mitochondrial morphology is irreversible, and these later changes tilt the scales toward mitochondrial fission, leading to mitochondrial fragmentation and reduced functionality [69]. Excess APAP-generated NAPQI interferes with the electron transport chain, when complex III is responsible for leaking electrons, which come into contact with oxygen and generate superoxide radicals [71]. The produced ROS activate the JNK pathway, and phosphorylated JNK is transferred to the mitochondria and binds to the Sab protein, further aggravating oxidative stress and ROS production [72]. Superoxide radicals are converted into hydrogen peroxide (H2O2) and molecular oxygen (O2−) by manganese superoxide dismutase (MnSOD) in the mitochondria or produce peroxynitrite (ONOO) when combined with NO. GSH or antioxidant enzymes (such as catalase) can scavenge H2O2. Excessive free radicals cause GSH depletion, which results in the accumulation of ONOO in the mitochondria, eventually leading to mitochondrial DNA damage and the formation of nitrotyrosine protein adducts. The mitochondrial damage caused by adduct formation triggers the collapse of the mitochondrial membrane potential under redox conditions and hepatotoxicity, resulting in mitochondrial stress and liver damage [73]. Active JNK induces changes in the mitochondrial membrane permeability, leading to mitochondrial swelling and necrosis, culminating in liver cell damage [70]. Mitochondrial oxidative stress initiates extracellular death signaling, causing extracellular proteins (such as Bax) to migrate to the mitochondria. Bax forms pores on the outer membrane of the mitochondria, leading to the release of intermembrane proteins [60]. These proteins are then transferred to the nucleus, resulting in DNA breakage. In addition, mitochondrial oxidative stress and peroxynitrite cause MPT (mitochondrial permeability transition) through the MPTP (mitochondrial permeability transition pore). MPT can trigger different cellular responses, from the physiological regulation of mitophagy to the activation of apoptosis or necrosis [74].
  • Isoniazid-induced mitochondrial stress
Although INH and APAP share pathogenic mechanisms in liver injury, the hepatotoxicity of INH is more severe than that of APAP [8]. In INH-induced DILI, isoniazid is catalyzed by CYP2E1, generating the toxic metabolites acetylhydrazine and hydrazine. The covalent binding of acetylhydrazine to biomacromolecules in hepatocytes causes liver injury. The former can hydrolyze into hepatotoxic hydrazine, leading to the depletion of GSH, resulting in changes in ERS and mitochondrial membrane permeability (Figure 1). Furthermore, hydrazine forms superoxide by inhibiting mitochondrial complex II, promoting mitochondrial stress and liver injury [75]. It can also directly cause mitochondrial damage owing to the formation of isoniazid or its metabolites. For example, INH reduces ATP production in mitochondria by blocking the electron flow, which promotes oxidative stress and energy homeostasis imbalance in the mitochondria [76].
  • Valproic acid induced mitochondrial stress
Valproic acid (VPA) is an anti-seizure drug that causes idiosyncratic liver injury [77]. The VPA metabolite, 2-propyl-4-pentenoic acid, has been implicated in VPA-induced hepatotoxicity [78]. Furthermore, VPA can deplete inositol, which increases the expression of fatty acid elongases, catalyzing the synthesis of fatty acids and ceramides. Increased ceramide levels decrease the expression of amino acid transporters, which induces cellular stress due to nutritional stress and ERS induction [21]. VPA and its metabolites interfere with mitochondrial function mainly through the upregulation of the fatty acid transporter (CD36) and inhibition of β-oxidation, which increases fatty acid content, decreases ATP and NADH, and inhibits the mitochondrial respiratory chain. This further increases ROS, leads to cytochrome c release, activates caspases, and results in mitochondrial stress, ultimately causing hepatocyte apoptosis and liver injury [41]. VPA also upregulates progesterone, which promotes an accumulation of cholesterol in the mitochondria and the structural alteration of the mitochondrial surface transporters, ultimately inducing mitochondrial stress [22].

3. ERS Triggers and Signaling Pathways in DILI

Several endogenous and exogenous factors can affect ER function and cause ERS. Exogenous factors include oxygen deficit, radiation, toxic chemicals, and infection by pathogenic microorganisms [57,79]. Endogenous factors include abnormal calcium regulation, lipid metabolism disorders, etc. They are described in detail below because DILI mainly involves endogenous factors.

3.1. ERS Triggers

3.1.1. Abnormal Calcium Regulation

The thapsigargin-induced ERS involves the inhibition of sarco/endoplasmic reticulum Ca2+-ATPase (SERCA), which results in severe calcium depletion in the ER. In addition to the typical inducers of ERS, the rise in free cytosolic calcium also induces apoptosis in different types of cells treated with thapsigargin, which induces ERS by inhibiting the SERCA, resulting in serious consumption of ER calcium [80]. In addition to the activation of downstream effectors that trigger cell death induced by ERS, an increase in free cytoplasmic calcium is also an effective pro-apoptotic signal in different types of cells. Trimetazidine, matrine, lonomycin, and other drugs can trigger ERS by disrupting normal calcium metabolism.

3.1.2. Lipid Metabolism Disorders

The liver plays a central role in lipid homeostasis; under normal physiological conditions, the lipid input is equal to the lipid output of the body. The interruption of this balance promotes lipid metabolism disorders. Rituximab, an HIV protease inhibitor (HPI), can cause lipid metabolism disorder by increasing the accumulation of free cholesterol that consumes the ER calcium pool [81], increases SREBP activity to activate UPR, induces apoptosis, and promotes the formation of foam cells [82].

3.2. Endoplasmic Reticulum Stress Signaling Pathways Involved in DILI

3.2.1. Three UPR Pathways

Misfolded proteins are prone to toxic aggregation; therefore, eukaryotes have evolved a UPR to ensure the normal dynamic balance of protein folding. The UPR maintains protein folding in the ER, thereby inhibiting the toxicity associated with the accumulation of unfolded proteins [83]. This process occurs under the regulation of immunoglobulin heavy chain binding protein (Bip, GRP78), an amino-terminal chaperone of the endoplasmic lumen end. An increase in unfolded protein promotes the dissociation of Bip, releasing its inhibitory effects over three transmembrane proteins and initiating the stress response. These three proteins, ATF6, inositol-requiring enzyme-1α (IRE1α), and PKR-like ER kinase (PERK), help the ER recover from stress [84].
First, activated ATF6 enters the Golgi and hydrolyzes the S1P and S2P (Sphingosine-1-phosphate) enzymes. Subsequently, ATF6 dissociates from the dictyosome and enters the nucleus as a transcription factor binding to the ERSE (ERS-response element), inducing the transcription of multiple genes, including Bip, CHOP, and X-box binding protein 1 (XBP1) (Figure 3) [85].
Second, the IRE1 pathway mediates the transcriptional induction of ER quality control (including molecular chaperones, folding enzymes, and endoplasmic reticulum-associated degradation components) and secretory proteins. IRE1α activates its nuclease activity via homodimerization and autophosphorylation [86]. IRE1α splices XBP1 mRNA to obtain the transcription factor XBP1, which is transferred into the nucleus to regulate UPR and ERAD-related genes and alleviate the damage caused by the ERS environment. However, sustained and severe ERS promotes apoptotic signaling through the IRE-1 pathway. Activated IRE1α can recruit TRAF2 (TNF receptor-associated factor 2) and with ASK1, phosphorylate JNK and NF-κB (Figure 4). JNK phosphorylates several Bcl-2 family members and promotes cytochrome c release, caspase activation, and apoptosis. Thus, signals initiated from the cytoplasmic kinase domain of IRE1α are primarily pro-apoptotic signals involving the JNK pathway [87]. The NF-κB pathway is involved in apoptosis, cell survival, and the activation of inflammatory cells. Continuous activation of this pathway can lead to uncontrolled cell growth and increased levels of tumor growth (such as IL-1β, TNF, and IL-6) and anti-apoptotic factors. This pathway can also activate autophagy in a MAPK-dependent manner, eventually leading to hepatic stellate cell (HSC) activation and liver fibrosis. However, ERS promotes the apoptosis of activated HSC, reducing and rescuing liver fibrosis [88]. Liver pathological manifestation of chronic liver diseases, such as non-alcoholic liver injury, chronic hepatitis B, and chronic hepatitis C. If not treated actively, damaged hepatocytes can stimulate the proliferation of intrahepatic fibrous connective tissue, progressing from liver fibrosis to cirrhosis.
Third, Bip release activates the PERK pathway that promotes the phosphorylation of eIF2α, triggering an integrated stress response (ISR) that inhibits protein translation and reduces protein load in the endoplasmic reticulum [89]. This phenomenon increases the expression of ATF4 (Activating Transcription Factor 4), which enters the nucleus to regulate the expression of UPR target genes. Under the condition of ERS transition, PERK promotes eIF2α phosphorylation and induces ATF4 translation expression through homologous dimerization and autophosphorylation self-activation. Eventually, the transcription of the ERS apoptosis marker protein CHOP is promoted, culminating in the activation of the apoptosis signaling pathway (Figure 5) [90].
Following ERS induction, the body restores cell homeostasis through the UPR, and the final fate of the cells is determined by several factors. You et al. found that the UPR response transcription factor, QRICH1, determines whether cells move towards adaptation or programmed death during the final UPR period. As a regulator of a unique transcription module, QRICH1 coordinates cellular stress responses to regulate protein synthesis and secretion under steady state and pathological conditions [91]. The consequences of ERS also depend on the duration and intensity of the stress [92].

3.2.2. Caspase-12 Pathway

After ERS, the caspase-12 pathway is activated. ERS induces high CHOP expression, which promotes hepatocyte apoptosis by downregulating the expression of the apoptosis gene Bcl-2 combined with the UPR-related molecules. ERS causes a calcium imbalance in the ER and activates caspase-12, which subsequently activates caspase-9 and caspase-3 by inducing an increase in molecular chaperones, such as Bip and GRP94, ultimately inducing hepatocyte apoptosis. Under steady-state conditions, TRAF2 forms a stable complex with caspase-12, and endoplasmic reticulum stress induces caspase-12 to separate from TRAF2 and promote its dimerization (Figure 6) [93].

4. Mitochondrial Stress Signaling Channels Activated by Drug-Induced Liver Injury

Integrated stress response (ISR) is an evolutionarily conserved intracellular signaling network activated in response to internal and external stresses. ISR is induced by four eIF2α kinases, namely PERK, amino acid general control non-inhibitory 2 (GCN2, which senses amino acid reduction), double-stranded RNA-dependent protein kinase (PKR, which senses viruses), and eIF2α kinase (HRI, which senses heme deficiency). UPR in the ER mainly includes the IRE1α, ATF6, and PERK pathways. The PERK-eIF2α-ATF4 pathway is a branch of the ISR; mitochondrial stress can cause eIF2α phosphorylation through ISR, inhibiting overall protein translation while upregulating the translation of proteins, such as ATF4 and ATF5, which initiate the mitochondrial UPR (UPRmt) [94]. Two mitochondrial stress conduction pathways exist in mammals: the ATF4-ATF5-CHOP and the OMA1–DELE1–HRI pathways [95].
The first mitochondrial stress signal channel is the ATF4-ATF5-CHOP. Mitochondrial dysfunction stimulates the phosphorylation of eIF2α, which translates to the transcription factors CHOP, ATF4, and ATF5. The ATF5 activity is negatively regulated by mitochondrial import into healthy mitochondria. Under steady-state conditions, ATF5 is introduced into mitochondria and degraded by LONP1. Under mitochondrial stress, CHOP, ATF4, and ATF5 are transferred to the nucleus, and UPRmt occurs. ISR stimulates the upstream open reading frame (uORF)-regulated translation of CHOP, ATF4, and ATF5 proteins, which subsequently activate transcription, reconnect cell metabolism, and enhance mitochondrial protein homeostasis. Mammals When the mitochondrial stress time exceeds a certain threshold, mitochondria are cleared through mitotic phagocytosis [83].
The second stress signal channel is the OMA1–DELE1–HRI pathway. Kampmann et al. found that mitochondrial stress is transmitted to the cytoplasm through the OMA1–DELE1–HRI pathway [96]. Mitochondrial stress stimulates OMA1 (a mitochondrial stress-activated protease)-dependent cleavage of DELE1, which promotes its accumulation in the cytosol in the form of DELE1s. The cytosolic DELE1s bind to HRI, activate eIF2α kinase activity, phosphorylate eIF2α, and upregulate downstream ATF4 translation, triggering ISR to ensure mitochondrial homeostasis [95]. Phosphorylation of eIF2α leads to a decrease in overall protein levels but an increase in the transcription factors ATF4, ATF5, and CHOP (Figure 7). Basic leucine zipper (bZIP) regulates the upregulation of transcription factor ATF4 translation that binds to DNA targets to promote cell survival and maintain homeostasis. However, when stress is severe or when its duration exceeds a certain threshold, the ISR signal is converted to regulate apoptosis and clear damaged mitochondria [97]. Additionally, Nour et al. found that HRI is essential for the downstream signaling of NOD1 and NOD2, which can induce NF-κB signaling [98], activate inflammatory cells, and cause inflammation in the liver.

5. Communication between Mitochondrial and Endoplasmic Reticulum Stress

5.1. Calcium Exchange

5.1.1. Contact Site between ER and Mitochondria-MAM

The ER and mitochondria are in physical contact through a dynamic membrane, the mitochondria-associated membrane (MAM). The MAM contains cholesterol and sphingolipids for increased thickness [99,100] and an intracellular lipid raft-like domain that is closely apposed to the mitochondria, both physically and biochemically. The ER and mitochondria undergo calcium ion exchange, energy exchange, and lipid transport through the MAM. MAM can maintain the steady state of both organelles through contact between the ER and mitochondrial membrane. Therefore, the operation of normal functions of the MAM plays an important role in ensuring the steady state of the human body. Studies have shown that MAM-localized proteins can regulate the UPR. PERK is a key controller of oxygen and phosphorus; although it is not located in the mitochondria, it can be found in the MAM. When the concentration of cytosolic Ca2+ increases, dimerized PERK interacts with silk protein A, rearranges the protein cytoskeleton, and increases the MC (mitochondrial complex) formation. This interaction promotes the formation of endoplasmic reticulum-plasma membrane contacts [101]. IRE1α has also been shown in MAM; furthermore, the interaction of IP3R3 determines the structure and the Ca2+ signaling on the MAM [102].

5.1.2. ER Receptor-Mediated Ca2+ Exchange

The sigma-1 receptor (Sig-1R) is a resident protein of the ER that localizes to the MAM. Deletion of Sig-1R can damage ER-mitochondrial contact sites and affect intracellular calcium signal transduction and mitochondrial morphology, accompanied by ERS and mitochondrial defects [103]. IP3R (inositol 1,4,5-triphosphate receptor) releases calcium from the endoplasmic reticulum and increases the intracellular calcium concentration. Sig-1R forms a complex to stabilize IP3R with the ER partner, Bip, on the MAM; however, when Ca2+ in the ER is depleted or stimulated by ligands, Sig-1R dissociates from Bip, leading to the long-term signal transduction of Ca2+ into the mitochondria through IP3R. Excessive ERS leads to an imbalance in protein homeostasis. The release of calcium ions from the ER into the cytoplasm leads to the opening of the mitochondrial permeability transition pores, forming a positive feedback regulation and bidirectional apoptosis [104]. As mentioned previously, calcium imbalance also activates caspase-12 and the production of pro-inflammatory cytokines, such as IL-1β and IL-18, a key step in ERS-induced apoptosis. Calcium imbalance also affects mitochondrial-induced apoptosis by coupling Bip on the MAM with voltage-dependent anion channel protein 1 (VDAC1) (Figure 8) [71].

5.1.3. TRPM2 Involved in Ca2+ Exchange

Transient receptor potential 2 (TRPM2) is a nonselective cation channel located on the cell membrane. Several intracellular signaling pathways mediate ROS-induced hepatocyte injury; however, one of the main events is the ROS-induced increase in Ca2+ influx and the subsequent increase in cytoplasmic Ca2+ concentration [105,106,107]. The TRPM2 channel allows Ca2+ and Na+ to flow through the plasma membrane and release Ca2+ and Zn2+ from lysosomes. In different animal cell types, calcium mediates ROS-induced cell damage and death through the TRPM2 channel. Under oxidative stress, ROS activates poly ADPR polymerase (PARP), which induces ADPR production and TRPM2 channel activation, increasing calcium influx [108]. TRPM2 transport from the intracellular membranes to the plasma membrane contributes to the activation of TRPM2 in APAP-induced hepatocytes [16]. The activation of the TRPM2/Ca2+/CaMKII signaling pathway promotes lipid accumulation, mitochondrial damage, and ERS, which aggravates the progression of non-alcoholic fatty liver disease (NAFLD) to non-alcoholic steatohepatitis (NASH), then to cirrhosis and ultimately, to liver cancer [109]. Overall, calcium homeostasis in intracellular mitochondria and the endoplasmic reticulum plays a determinant role in maintaining human health.

5.2. Relationship between Mitochondrial and Endoplasmic Reticulum Stress

Xiao et al. found that chemical-induced ERS can stimulate the expression of the mitochondrial stress protein HSP60, impair mitochondrial respiration, and reduce mitochondrial membrane potential in mouse hepatocytes. Moreover, HSP60 overexpression induces ERS by increasing Bip and CHOP levels. HSP60 regulates ERS-induced liver lipid production through the mTORC1-SREBP1 signaling pathway [13]. ERS can cause eIF2α phosphorylation through ISR, inhibit protein translation, and upregulate the translation of proteins such as ATF4 to initiate mitochondrial stress [83,89]. These reports suggest that mitochondrial stress is a consequence of endoplasmic reticulum stress.
Research shows that a large amount of ROS is produced upon mitochondrial stress, and when the redox balance in the endoplasmic reticulum is disrupted, which triggers ERS [72]. Therefore, it appears that ERS is also a consequence of mitochondrial dysfunction. ERO1α-PDI (protein disulfide isomerase) in ER catalyzes the formation of disulfide bonds in proteins, transferring electrons to oxygen to produce H2O2 [110]. This produces a large amount of ROS and releases a large amount of Ca2+ into the cytoplasm, which results in the opening of the MPTP and mitochondrial swelling, thereby aggravating mitochondrial oxidative stress. Therefore, swelling of mitochondria and ERS do not exist independently in severe DILI, mutually affecting each other through various channels to determine liver cell damage.

6. The Immunological Mechanisms of APAP-Induced Liver Injury

In the immune response induced by hepatotoxicity-mediated APAP liver injury, Kupffer cells form the first line of defense by recognizing necrotic hepatocytes through danger-associated molecular pattern molecules (DAMPs) and pathogen-associated molecular patterns (PAMPs). Kupffer cells release cytokines and chemokines upon activation, leading to acute inflammation marked by an increase in acute phase proteins [111]. Concomitantly, cytokines amplify the inflammatory process by promoting the release of other inflammatory mediators from infiltrating leukocytes, which upregulate adhesion molecules (for example, ICAM-1 and CD11b/CD18), regulating immune cell aggregation through mediator production. The expression of adhesion molecules also contributes to neutrophil adhesion and translocation within the sinusoids and dependent oxidative stress that promotes hepatocyte death [112,113]. Additionally, activated B cells produce antibodies participating in APAP-induced liver injury. Although several studies have shown that the immune response plays an important role in DILI [1,43,114], its association with ERS and mitochondrial stress requires investigation.

7. Pathways for the Treatment of a Drug-Induced Liver Injury

7.1. SPHK1–STARD1 Pathway

SPHK1 deficiency leads to a decrease in S1P content, inhibiting MPT, ATF4, and ATF6 levels, and inflammatory gene expression. SPHK1 plays an important role in hepatotoxicity induced by ERS and mitochondrial stress. Using SPHK1 antagonists and gene knockouts in mice treated with APAP, Li et al. found that ERS was alleviated in mice, MPT was inhibited, expression of inflammatory factors was suppressed, and the risk of hepatotoxicity was reduced [14]. Therefore, targeting SPHK1 may be a novel strategy for the treatment of drug hepatotoxicity.
In mouse studies, Torres et al. found that after ERS occurs, the body mediates the upregulation of STARD1 through the phosphorylation of SH3BP5, JNK1, and JNK2 [67], which results in increased mitochondrial cholesterol content and hepatotoxicity. Knockout of STARD1 can prevent cholesterol accumulation in mitochondria, thereby reducing the risk of hepatotoxicity [15]. Therefore, targeting STARD1 may provide a novel strategy for hepatotoxicity treatment, but effective inhibitors are yet to be developed.
Knockout of SPHK1 alleviates mitochondrial and ERS, whereas the knockout of STARD1 only can alleviate ERS. Inhibition of the SPHK1–STARD1 pathway can reduce DILI through the alleviation of organelle stress. Whether the knockout of SPHK1 affects STARD1 or the knockout of STARD1 affects SPHK1 expression needs to be determined to identify the upstream trigger of the SPHK1–STARD1 pathway. Therefore, studying the mechanism of this pathway and its inhibitors is important as it might prove to be a novel strategy for treating drug-induced hepatotoxicity.

7.2. ATF4-ATF5-CHOP Pathway

ATF4 is the main mediator of ISR; Koumenis et al. showed ATF4 upregulation in every tumor studied [115]. After ATF4 knockout, tumor cells die from an excessive internal stress response [17] and exhibit reduced angiogenesis without causing significant damage to mice. The Koumenis Lab is currently working on the development of ATF4 inhibitors. Considering its role in DILI pathogenesis, ATF4 is a potential target for DILI treatment.
ATF5 has tumor cell type heterogeneity and plays a role in promoting tumorigenesis and deterioration in several tumors. ATF5 expression is reduced in liver cancer tissues and acts as a tumor suppressor gene [116]. CHOP is a marker of ERS, and its inhibition can effectively reduce the consequences of ERS [117]. The commonly used methods to investigate this pathway include siRNA-targeted silencing and shRNA-targeted interference, CHOP antibodies, and CHOP inhibitors. Both ATF5 and CHOP might be pivotal targets for DILI therapeutic approaches.

7.3. TRPM2 Pathway

Zhang et al. found that the knockdown of TRPM2 alleviated injury through the activation of autophagy and inhibition of the NLRP3 inflammasome pathway [118]. In TRPM2 knockout mice, the APAP-induced liver injury, assessed through the blood concentration of liver enzymes and histology, was significantly decreased compared with that in wild-type mice [119]. This finding suggests TRPM2 is a therapeutic target for APAP-induced liver injury.
Curcumin is a natural antioxidant product that prevents ROS-induced liver injury by inhibition of TRPM2. Even at nanomolar concentrations, curcumin inhibits Ca2+ entry into hepatocytes through the TRPM2 channels [120]. This phenomenon shows that the TRPM2 channel is a potential pharmacological target for the prevention of ROS-induced liver injury. Professor Zhang Liang Ren from Peking University developed a class of TRPM2 channel inhibitors that are structural derivatives of anthranilic acid. Broadly speaking, A23 is a highly active and selective TRPM2 inhibitor [121].

7.4. JNK Signaling Pathway

The processing and folding of proteins in the endoplasmic reticulum cause ERS, which activates JNK. There are several compounds used to inhibit JNK signaling. The most used is SP600125 [122], which can reduce liver mitochondrial oxidized glutathione (GSSG) levels while promoting liver GSH recovery, revealing a protective effect mainly due to the inhibition of mitochondrial oxidative stress [123]. Leflunomide (LEF) is a commonly used antirheumatic drug in clinical practice that protects hepatocytes from APAP-induced liver injury. LEF can inhibit the phosphorylation of JNK1 and JNK2, preventing changes in mitochondrial permeability and the release of pro-apoptotic factors, thereby protecting liver cells from damage [124]. BI-78D3 is another JNK inhibitor that effectively prevents CCl4-induced acute liver injury [125]. Additionally, ASK1 is a MAP3 kinase that activates the downstream terminal kinases, JNK and p38. The ASK1 inhibitor, GS-444217, ameliorated NASH and improved fibrosis in preclinical studies and a short-term clinical trial [126]; however, this inhibitor was not effective in late-phase clinical trials [127,128]. Currently, biologics and the transplantation of stem cells have been shown to inhibit JNK in DILI through modulation of GSH resynthesis [129].

7.5. NF-κB Pathway

Intracellular and extracellular stimuli bind tumor necrosis factor receptor-associated factor (TRAF) to the receptor interacting protein (RIP). RIP proximal signal adaptor protein activates the IKB kinase complex (IKK). IKK activates IKB kinase, which phosphorylates and ubiquitinates IKB protein, degrades, and releases the NF-κB dimer. Finally, NF-κB is activated, translocates to the nucleus, and promotes transcription of target genes. NF-κB is a central link between liver injury, fibrosis, and hepatocellular carcinoma and represents a target for the prevention or treatment of DILIs [130]. Some common inhibitors associated with NF-κB channels are listed in Table 2.

7.6. UPR Pathway

Here we describe that PERK, IRE1, and ATF6 can induce apoptosis and activate inflammatory pathways by mediating the phosphorylation of eIF2α, which can cause different degrees of liver damage. The stress response can be reduced by inhibiting the expression of these three proteins. Several available compounds are currently used to regulate UPR, but some challenges and limitations to their use remain. In addition, harm can be reduced by inducing the generation of Bip. The main inhibitors are listed in Table 2. In addition to inhibitors, the CRISPR/Cas9 system can also be used for UPR pathway gene knockout.

8. Conclusions

DILI is a complex disease modulated by numerous metabolic, genetic, and environmental mechanisms. Mitochondrial stress and ERS are key factors in DILI, such that liver injury is both a cause and a consequence of these events, creating a positive feedback loop that may promote the development and progression of hepatic injury. Here, we propose that the SPHK1–STARD1 pathway may play a role in determining the outcomes of drug-induced liver damage. This concept requires further experimental testing to identify which other targets might be used to therapeutically prevent or delay DILI progression. Over the last decade, a significant effort has been invested in targeting signaling proteins involved in organelle stress and an array of inhibitors is now available to be tested in the clinic. However, these molecules show limitations, off-target effects being a case in point. A further challenge to be addressed in the future will be to develop tool compounds into agents that can safely be administered in the clinic.

Author Contributions

Conceptualization, M.W. and Y.P.; investigation, Q.Z., T.Y. (Ting You), T.Y. (Tao Yue) and Y.Z.; writing—original draft preparation, S.P.; writing—review and editing, M.W., Y.P., Q.Z. and S.P.; visualization, S.P.; funding acquisition, M.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by [National Natural Science Foundation of China] grant number [32160850], [Gansu Natural Science Foundation of China] grant number [22JR5RA868] and [Key Talent Project of Gansu Province] grant number [2022-0623-RCC-0307].

Data Availability Statement

All data presented in this study are available on request from the corresponding authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Björnsson, H.; Björnsson, E. Drug-induced liver injury: Pathogenesis, epidemiology, clinical features, and practical management. Eur. J. Intern. Med. 2022, 97, 26–31. [Google Scholar] [CrossRef] [PubMed]
  2. Reuben, A.; Koch, D.G.; Lee, W.M. Drug-induced acute liver failure: Results of a U.S. multicenter, prospective study. Hepatology 2010, 52, 2065–2076. [Google Scholar] [CrossRef] [Green Version]
  3. Ou, P.; Liu, X.; Tang, Z.; Hou, Z.; Liu, L.; Liu, J.; Zhou, S.; Fang, Z.; Sun, K.; Chen, Y.; et al. Gynura Segetum Related Hepatic Sinusoidal Obstruction Syndrome: A Liver Disease with High Mortality and Misdiagnosis Rate. Curr. Pharm. Des. 2019, 25, 3762–3768. [Google Scholar] [CrossRef] [PubMed]
  4. Iorga, A.; Dara, L. Cell death in drug-induced liver injury. Adv. Pharmacol. 2019, 85, 31–74. [Google Scholar] [CrossRef] [PubMed]
  5. Chang, L.; Xu, D.; Zhu, J.; Ge, G.; Kong, X.; Zhou, Y. Herbal Therapy for the Treatment of Acetaminophen-Associated Liver Injury: Recent Advances and Future Perspectives. Front. Pharmacol. 2020, 11, 313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Su, Q.; Kuang, W.; Hao, W.; Liang, J.; Wu, L.; Tang, C.; Wang, Y.; Liu, T. Antituberculosis Drugs (Rifampicin and Isoniazid) Induce Liver Injury by Regulating NLRP3 Inflammasomes. Mediat. Inflamm. 2021, 2021, 8086253. [Google Scholar] [CrossRef]
  7. Zhao, H.; Wang, Y.; Zhang, T.; Wang, Q.; Xie, W. Drug-Induced Liver Injury from Anti-Tuberculosis Treatment: A Retrospective Cohort Study. Med. Sci. Monit. 2020, 26, e920350. [Google Scholar] [CrossRef]
  8. Wang, N.; Wang, H.; Zhang, J.; Ji, X.; Su, H.; Liu, J.; Wang, J.; Zhao, W. Endogenous peroxynitrite activated fluorescent probe for revealing anti-tuberculosis drug induced hepatotoxicity. Chin. Chem. Lett. 2022, 33, 1584–1588. [Google Scholar] [CrossRef]
  9. Marciniak, S.J.; Chambers, J.E.; Ron, D. Pharmacological targeting of endoplasmic reticulum stress in disease. Nat. Rev. Drug Discov. 2022, 21, 115–140. [Google Scholar] [CrossRef]
  10. Boelsterli, U.A.; Lim, P.L. Mitochondrial abnormalities—A link to idiosyncratic drug hepatotoxicity? Toxicol. Appl. Pharmacol. 2007, 220, 92–107. [Google Scholar] [CrossRef]
  11. Jaeschke, H.; McGill, M.R.; Ramachandran, A. Oxidant stress, mitochondria, and cell death mechanisms in drug-induced liver injury: Lessons learned from acetaminophen hepatotoxicity. Drug Metab. Rev. 2012, 44, 88–106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Marchi, S.; Patergnani, S.; Missiroli, S.; Morciano, G.; Rimessi, A.; Wieckowski, M.R.; Giorgi, C.; Pinton, P. Mitochondrial and endoplasmic reticulum calcium homeostasis and cell death. Cell Calcium 2018, 69, 62–72. [Google Scholar] [CrossRef] [PubMed]
  13. Xiao, T.; Liang, X.; Liu, H.; Zhang, F.; Meng, W.; Hu, F. Mitochondrial stress protein HSP60 regulates ER stress-induced hepatic lipogenesis. J. Mol. Endocrinol. 2020, 64, 67–75. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Li, L.; Wang, H.; Zhang, J.; Sha, Y.; Wu, F.; Wen, S.; He, L.; Sheng, L.; You, Q.; Shi, M.; et al. SPHK1 deficiency protects mice from acetaminophen-induced ER stress and mitochondrial permeability transition. Cell Death Differ. 2020, 27, 1924–1937. [Google Scholar] [CrossRef]
  15. Torres, S.; Solsona-Vilarrasa, E.; Nuñez, S.; Matías, N.; Insausti-Urkia, N.; Castro, F.; Casasempere, M.; Fabriás, G.; Casas, J.; Enrich, C.; et al. Acid ceramidase improves mitochondrial function and oxidative stress in Niemann-Pick type C disease by repressing STARD1 expression and mitochondrial cholesterol accumulation. Redox Biol. 2021, 45, 102052. [Google Scholar] [CrossRef] [PubMed]
  16. Kheradpezhouh, E.; Zhou, F.H.; Barritt, G.J.; Rychkov, G.Y. Oxidative stress promotes redistribution of TRPM2 channels to the plasma membrane in hepatocytes. Biochem. Biophys. Res. Commun. 2018, 503, 1891–1896. [Google Scholar] [CrossRef]
  17. Tameire, F.; Verginadis, I.I.; Leli, N.M.; Polte, C.; Conn, C.S.; Ojha, R.; Salinas, C.S.; Chinga, F.; Monroy, A.M.; Fu, W.; et al. ATF4 couples MYC-dependent translational activity to bioenergetic demands during tumour progression. Nat. Cell Biol. 2019, 21, 889–899. [Google Scholar] [CrossRef]
  18. Wang, T.; Zhao, X.; Shao, C.; Ye, L.; Guo, J.; Peng, N.; Zhang, H.; Li, J.; Kong, Y.; You, H.; et al. A proposed pathologic sub-classification of drug-induced liver injury. Hepatol. Int. 2019, 13, 339–351. [Google Scholar] [CrossRef]
  19. Hoofnagle, J.H.; Björnsson, E.S. Drug-Induced Liver Injury—Types and Phenotypes. N. Engl. J. Med. 2019, 381, 264–273. [Google Scholar] [CrossRef]
  20. Björnsson, E.; Talwalkar, J.; Treeprasertsuk, S.; Kamath, P.S.; Takahashi, N.; Sanderson, S.; Neuhauser, M.; Lindor, K. Drug-induced autoimmune hepatitis: Clinical characteristics and prognosis. Hepatology 2010, 51, 2040–2048. [Google Scholar] [CrossRef]
  21. Jadhav, S.; Russo, S.; Cottier, S.; Schneiter, R.; Cowart, A.; Greenberg, M.L. Valproate Induces the Unfolded Protein Response by Increasing Ceramide Levels. J. Biol. Chem. 2016, 291, 22253–22261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Li, S.; Qi, J.; Sun, Y.; Gao, X.; Ma, J.; Zhao, S. An integrated RNA-Seq and network study reveals that valproate inhibited progesterone production in human granulosa cells. J. Steroid Biochem. Mol. Biol. 2021, 214, 105991. [Google Scholar] [CrossRef] [PubMed]
  23. Munz, M.; Grummich, H.; Birkmann, J.; Wilhelm, M.; Holzgrabe, U.; Sörgel, F. Severe Drug-Induced Liver Injury as an Adverse Drug Event of Antibiotics: A Case Report and Review of the Literature. Chemotherapy 2017, 62, 367–373. [Google Scholar] [CrossRef] [PubMed]
  24. Hou, L.; Liu, F.; Zhao, C.; Fan, L.; Hu, H.; Yin, S. Combination of oxytetracycline and quinocetone synergistically induces hepatotoxicity via generation of reactive oxygen species and activation of mitochondrial pathway. Toxicol. Mech. Methods 2022, 32, 49–57. [Google Scholar] [CrossRef]
  25. Björnsson, E.S. Drug-induced liver injury due to antibiotics. Scand. J. Gastroenterol. 2017, 52, 617–623. [Google Scholar] [CrossRef]
  26. Gupta, A.; Singh, A.K.; Faridi, K.; Jain, P. Cefazolin Induced Liver Injury and Hypoprothrombinemia. J. Clin. Exp. Hepatol. 2018, 8, 213–214. [Google Scholar] [CrossRef]
  27. Zoubek, M.E.; Lucena, M.I.; Andrade, R.J.; Stephens, C. Systematic review: Ibuprofen-induced liver injury. Aliment. Pharmacol. Ther. 2020, 51, 603–611. [Google Scholar] [CrossRef]
  28. Wen, C.; Zhuang, Z.; Song, H.; Tong, S.; Wang, X.; Lin, Y.; Zhan, H.; Chen, Z.; Hu, L. Metabolism of liver CYP450 and ultrastructural changes after long-term administration of aspirin and ibuprofen. Biomed. Pharmacother. 2018, 108, 208–215. [Google Scholar] [CrossRef]
  29. Jung, S.-H.; Lee, W.; Park, S.-H.; Lee, K.-Y.; Choi, Y.-J.; Choi, S.; Kang, D.; Kim, S.; Chang, T.-S.; Hong, S.-S.; et al. Diclofenac impairs autophagic flux via oxidative stress and lysosomal dysfunction: Implications for hepatotoxicity. Redox Biol. 2020, 37, 101751. [Google Scholar] [CrossRef]
  30. Gunther, M.; Dopheide, J.A. Antipsychotic Safety in Liver Disease: A Narrative Review and Practical Guide for the Clinician. J. Acad. Consult.-Liaison Psychiatry 2023, 64, 73–82. [Google Scholar] [CrossRef]
  31. Vukotić, N.T.; Đorđević, J.; Pejić, S.; Đorđević, N.; Pajović, S.B. Antidepressants- and antipsychotics-induced hepatotoxicity. Arch. Toxicol. 2021, 95, 767–789. [Google Scholar] [CrossRef]
  32. Shah, J.; Muir, J.; Furfaro, D.; Beitler, J.R.; Dzierba, A.L. Use of N-Acetylcysteine for Clozapine-Induced Acute Liver Injury: A Case Report and Literature Review. J. Pharm. Pract. 2021, 8971900211034007. [Google Scholar] [CrossRef] [PubMed]
  33. Björnsson, E.S. Hepatotoxicity of statins and other lipid-lowering agents. Liver Int. 2017, 37, 173–178. [Google Scholar] [CrossRef] [PubMed]
  34. Meunier, L.; Larrey, D. Chemotherapy-associated steatohepatitis. Ann. Hepatol. 2020, 19, 597–601. [Google Scholar] [CrossRef] [PubMed]
  35. Alessandrino, F.; Qin, L.; Cruz, G.; Sahu, S.; Rosenthal, M.H.; Meyerhardt, J.A.; Shinagare, A.B. 5-Fluorouracil induced liver toxicity in patients with colorectal cancer: Role of computed tomography texture analysis as a potential biomarker. Abdom. Radiol. 2019, 44, 3099–3106. [Google Scholar] [CrossRef]
  36. Honda, S.; Tsujimoto, M.; Minegaki, T.; Mori, T.; Muraoka, J.; Nishiguchi, K. A case of idiosyncratic liver injury after oxaliplatin-induced thrombocytopenia. J. Clin. Pharm. Ther. 2020, 45, 373–375. [Google Scholar] [CrossRef]
  37. Kim, J.-H.; Nam, W.S.; Kim, S.J.; Kwon, O.K.; Seung, E.J.; Jo, J.J.; Shresha, R.; Lee, T.H.; Jeon, T.W.; Ki, S.H.; et al. Mechanism Investigation of Rifampicin-Induced Liver Injury Using Comparative Toxicoproteomics in Mice. Int. J. Mol. Sci. 2017, 18, 1417. [Google Scholar] [CrossRef] [Green Version]
  38. Xu, Y.; Jiang, Y.; Li, Y. Pyrazinamide enhances lipid peroxidation and antioxidant levels to induce liver injury in rat models through PI3k/Akt inhibition. Toxicol. Res. 2020, 9, 149–157. [Google Scholar] [CrossRef]
  39. Wang, Y.-C.; Chen, K.-H.; Chen, Y.-L.; Lin, S.-W.; Liu, W.-D.M.; Wang, J.-T.M.; Hung, C.-C.M. Pyrazinamide related prolonged drug-induced liver injury: A case report. Medicine 2022, 101, e30955. [Google Scholar] [CrossRef]
  40. Chalasani, N.; Bonkovsky, H.L.; Stine, J.G.; Gu, J.; Barnhart, H.; Jacobsen, E.; Björnsson, E.; Fontana, R.J.; Kleiner, D.E.; Hoofnagle, J.H. Clinical characteristics of antiepileptic-induced liver injury in patients from the DILIN prospective study. J. Hepatol. 2022, 76, 832–840. [Google Scholar] [CrossRef]
  41. Ezhilarasan, D.; Mani, U. Valproic acid induced liver injury: An insight into molecular toxicological mechanism. Environ. Toxicol. Pharmacol. 2022, 95, 103967. [Google Scholar] [CrossRef] [PubMed]
  42. Guo, H.-L.; Jing, X.; Sun, J.-Y.; Hu, Y.-H.; Xu, Z.-J.; Ni, M.-M.; Chen, F.; Lu, X.-P.; Qiu, J.-C.; Wang, T. Valproic Acid and the Liver Injury in Patients with Epilepsy: An Update. Curr. Pharm. Des. 2019, 25, 343–351. [Google Scholar] [CrossRef] [PubMed]
  43. Rao, T.; Liu, Y.-T.; Zeng, X.-C.; Li, C.-P.; Ou-Yang, D.-S. The hepatotoxicity of Polygonum multiflorum: The emerging role of the immune-mediated liver injury. Acta Pharmacol. Sin. 2021, 42, 27–35. [Google Scholar] [CrossRef]
  44. Li, H.; Peng, Y.; Zheng, J. Dioscorea bulbifera L.-induced hepatotoxicity and involvement of metabolic activation of furanoterpenoids. Drug Metab. Rev. 2020, 52, 568–584. [Google Scholar] [CrossRef] [PubMed]
  45. Fan, W.; Fan, L.; Peng, C.; Zhang, Q.; Wang, L.; Li, L.; Wang, J.; Zhang, D.; Peng, W.; Wu, C. Traditional Uses, Botany, Phytochemistry, Pharmacology, Pharmacokinetics and Toxicology of Xanthium strumarium L.: A Review. Molecules 2019, 24, 359. [Google Scholar] [CrossRef] [Green Version]
  46. Lambert, A.; Cordeanu, M.; Gaertner, S.; Nouri, S.; Alt, M.; Stephan, D. Rivaroxaban-induced liver injury: Results from a venous thromboembolism registry. Int. J. Cardiol. 2015, 191, 265–266. [Google Scholar] [CrossRef] [PubMed]
  47. Licata, A.; Puccia, F.; Lombardo, V.; Serruto, A.; Minissale, M.G.; Morreale, I.; Giannitrapani, L.; Soresi, M.; Montalto, G.; Almasio, P.L. Rivaroxaban-induced hepatotoxicity: Review of the literature and report of new cases. Eur. J. Gastroenterol. Hepatol. 2018, 30, 226–232. [Google Scholar] [CrossRef]
  48. Song, A.B.; Rosovsky, R.P.; Connors, J.M.; Al-Samkari, H. Direct oral anticoagulants for treatment and prevention of venous thromboembolism in cancer patients. Vasc. Health Risk Manag. 2019, 15, 175–186. [Google Scholar] [CrossRef] [Green Version]
  49. Machlab, S.; Miquel, M.; Vergara, M.; Escoda, M.R.; Casas, M. Apixaban-induced liver injury. Rev. Esp. Enfermadades Dig. 2018, 111, 161–163. [Google Scholar] [CrossRef]
  50. Saha, M.; Sikder, P.; Saha, A.; Shah, S.; Sultana, S.; Emran, T.; Banik, A.; Islam, Z.; Islam, M.S.; Sharker, M.S.; et al. QbD Approach towards Robust Design Space for Flutamide/PiperineSelf-Emulsifying Drug Delivery System with Reduced Liver Injury. AAPS PharmSciTech 2022, 23, 62. [Google Scholar] [CrossRef]
  51. Thole, Z.; Salgueiro-Vázquez, E.; Revuelta, P.; Manso, G.; Hidalgo, A. Hepatotoxicity Induced by Antiandrogens: A Review of the Literature. Urol. Int. 2004, 73, 289–295. [Google Scholar] [CrossRef]
  52. Greenblatt, D.J.; Mikus, G. Ketoconazole and Liver Injury: A Five-Year Update. Clin. Pharmacol. Drug Dev. 2019, 8, 6–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Ogimura, E.; Nakagawa, T.; Deguchi, J.; Sekine, S.; Ito, K.; Bando, K. Troglitazone Inhibits Bile Acid Amidation: A Possible Risk Factor for Liver Injury. Toxicol. Sci. 2017, 158, 347–355. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Thomas, B.; Mohamed, M.; Alhail, M.; Awwad, F.A.; Wahba, R.M.; Hassan, S.B.; Omar, K.; El Kassem, W.; Rouf, P.A. A case of probable esomeprazole-induced transient liver injury in a pregnant woman with hyperemesis. Clin. Pharmacol. Adv. Appl. 2016, 8, 199–202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Ikemura, K.; Nakagawa, E.; Kurata, T.; Iwamoto, T.; Okuda, M. Altered Pharmacokinetics of Cimetidine Caused by Down-regulation of Renal Rat Organic Cation Transporter 2 (rOCT2) after Liver Ischemia-Reperfusion Injury. Drug Metab. Pharmacokinet. 2013, 28, 504–509. [Google Scholar] [CrossRef] [Green Version]
  56. Schwarz, D.S.; Blower, M.D. The endoplasmic reticulum: Structure, function and response to cellular signaling. Cell. Mol. Life Sci. 2016, 73, 79–94. [Google Scholar] [CrossRef] [Green Version]
  57. Almanza, A.; Carlesso, A.; Chintha, C.; Creedican, S.; Doultsinos, D.; Leuzzi, B.; Luís, A.; McCarthy, N.; Montibeller, L.; More, S.; et al. Endoplasmic reticulum stress signalling—From basic mechanisms to clinical applications. FEBS J. 2019, 286, 241–278. [Google Scholar] [CrossRef] [Green Version]
  58. Hwang, J.; Qi, L. Quality Control in the Endoplasmic Reticulum: Crosstalk between ERAD and UPR pathways. Trends Biochem. Sci. 2018, 43, 593–605. [Google Scholar] [CrossRef]
  59. Senft, D.; Ronai, Z.A. UPR, autophagy, and mitochondria crosstalk underlies the ER stress response. Trends Biochem. Sci. 2015, 40, 141–148. [Google Scholar] [CrossRef] [Green Version]
  60. Chao, X.; Wang, H.; Jaeschke, H.; Ding, W.-X. Role and mechanisms of autophagy in acetaminophen-induced liver injury. Liver Int. 2018, 38, 1363–1374. [Google Scholar] [CrossRef] [Green Version]
  61. Ishitsuka, Y.; Kondo, Y.; Kadowaki, D. Toxicological Property of Acetaminophen: The Dark Side of a Safe Antipyretic/Analgesic Drug? Biol. Pharm. Bull. 2020, 43, 195–206. [Google Scholar] [CrossRef] [Green Version]
  62. Oyadomari, S.; Mori, M. Roles of CHOP/GADD153 in endoplasmic reticulum stress. Cell Death Differ. 2004, 11, 381–389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Schröder, M.; Kaufman, R.J. The mammalian unfolded protein response. Annu. Rev. Biochem. 2005, 74, 739–789. [Google Scholar] [CrossRef]
  64. Foufelle, F.; Fromenty, B. Role of endoplasmic reticulum stress in drug-induced toxicity. Pharmacol. Res. Perspect. 2016, 4, e00211. [Google Scholar] [CrossRef] [PubMed]
  65. Nagy, G.; Kardon, T.; Wunderlich, L.; Szarka, A.; Kiss, A.; Schaff, Z.; Bánhegyi, G.; Mandl, J. Acetaminophen induces ER dependent signaling in mouse liver. Arch. Biochem. Biophys. 2007, 459, 273–279. [Google Scholar] [CrossRef] [PubMed]
  66. Kusama, H.; Kon, K.; Ikejima, K.; Arai, K.; Aoyama, T.; Uchiyama, A.; Yamashina, S.; Watanabe, S. Sodium 4-phenylbutyric acid prevents murine acetaminophen hepatotoxicity by minimizing endoplasmic reticulum stress. J. Gastroenterol. 2017, 52, 611–622. [Google Scholar] [CrossRef]
  67. Torres, S.; Baulies, A.; Insausti-Urkia, N.; Alarcón-Vila, C.; Fucho, R.; Solsona-Vilarrasa, E.; Núñez, S.T.; Robles, D.; Ribas, V.; Wakefield, L.; et al. Endoplasmic Reticulum Stress-Induced Upregulation of STARD1 Promotes Acetaminophen-Induced Acute Liver Failure. Gastroenterology 2019, 157, 552–568. [Google Scholar] [CrossRef] [Green Version]
  68. Nunnari, J.; Suomalainen, A. Mitochondria: In Sickness and in Health. Cell 2012, 148, 1145–1159. [Google Scholar] [CrossRef] [Green Version]
  69. Umbaugh, D.S.; Nguyen, N.T.; Jaeschke, H.; Ramachandran, A. Mitochondrial Membrane Potential Drives Early Change in Mitochondrial Morphology After Acetaminophen Exposure. Toxicol. Sci. 2021, 180, 186–195. [Google Scholar] [CrossRef]
  70. Han, D.; Dara, L.; Win, S.; Than, T.A.; Yuan, L.; Abbasi, S.Q.; Liu, Z.-X.; Kaplowitz, N. Regulation of drug-induced liver injury by signal transduction pathways: Critical role of mitochondria. Trends Pharmacol. Sci. 2013, 34, 243–253. [Google Scholar] [CrossRef] [Green Version]
  71. Nguyen, N.T.; Du, K.; Akakpo, J.Y.; Umbaugh, D.S.; Jaeschke, H.; Ramachandran, A. Mitochondrial protein adduct and superoxide generation are prerequisites for early activation of c-jun N-terminal kinase within the cytosol after an acetaminophen overdose in mice. Toxicol. Lett. 2021, 338, 21–31. [Google Scholar] [CrossRef] [PubMed]
  72. Win, S.; Than, T.A.; Min, R.W.M.; Aghajan, M.; Kaplowitz, N. c-Jun N-terminal kinase mediates mouse liver injury through a novel Sab (SH3BP5)-dependent pathway leading to inactivation of intramitochondrial Src. Hepatology 2016, 63, 1987–2003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Ramachandran, A.; Jaeschke, H. Acetaminophen Toxicity: Novel Insights Into Mechanisms and Future Perspectives. Gene Expr. 2018, 18, 19–30. [Google Scholar] [CrossRef] [PubMed]
  74. Mironova, G.D.; Pavlov, E.V. Mitochondrial Cyclosporine A-Independent Palmitate/Ca(2+)-Induced Permeability Transition Pore (PA-mPT Pore) and Its Role in Mitochondrial Function and Protection against Calcium Overload and Glutamate Toxicity. Cells 2021, 10, 125. [Google Scholar] [CrossRef]
  75. Vanova, K.H.; Kraus, M.; Neuzil, J.; Rohlena, J. Mitochondrial complex II and reactive oxygen species in disease and therapy. Redox Rep. 2020, 25, 26–32. [Google Scholar] [CrossRef] [Green Version]
  76. Chen, M.; Suzuki, A.; Borlak, J.; Andrade, R.J.; Lucena, M.I. Drug-induced liver injury: Interactions between drug properties and host factors. J. Hepatol. 2015, 63, 503–514. [Google Scholar] [CrossRef] [Green Version]
  77. Gheena, S.; Ezhilarasan, D.; Harini, K.S.; Rajeshkumar, S. Syringic acid and silymarin concurrent administration inhibits sodium valproate-induced liver injury in rats. Environ. Toxicol. 2022, 37, 2143–2152. [Google Scholar] [CrossRef]
  78. Mishra, M.K.; Kukal, S.; Paul, P.R.; Bora, S.; Singh, A.; Kukreti, S.; Saso, L.; Muthusamy, K.; Hasija, Y.; Kukreti, R. Insights into Structural Modifications of Valproic Acid and Their Pharmacological Profile. Molecules 2021, 27, 104. [Google Scholar] [CrossRef]
  79. Oakes, S.A.; Papa, F.R. The Role of Endoplasmic Reticulum Stress in Human Pathology. Annu. Rev. Pathol. Mech. Dis. 2015, 10, 173–194. [Google Scholar] [CrossRef] [Green Version]
  80. Zhou, H.-Y.; Sun, Y.-Y.; Chang, P.; Huang, H.-C. Curcumin Inhibits Cell Damage and Apoptosis Caused by Thapsigargin-Induced Endoplasmic Reticulum Stress Involving the Recovery of Mitochondrial Function Mediated by Mitofusin-2. Neurotox. Res. 2022, 40, 449–460. [Google Scholar] [CrossRef]
  81. Zhou, H.; Gurley, E.C.; Jarujaron, S.; Ding, H.; Fang, Y.; Xu, Z.; Pandak, W.M.; Hylemon, P.B. HIV protease inhibitors activate the unfolded protein response and disrupt lipid metabolism in primary hepatocytes. Am. J. Physiol.-Gastrointest. Liver Physiol. 2006, 291, G1071–G1080. [Google Scholar] [CrossRef] [PubMed]
  82. Zhou, H.; Pandak, W.M.; Lyall, V.; Natarajan, R.; Hylemon, P.B. HIV Protease Inhibitors Activate the Unfolded Protein Response in Macrophages: Implication for Atherosclerosis and Cardiovascular Disease. Mol. Pharmacol. 2005, 68, 690–700. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Smyrnias, I. The mitochondrial unfolded protein response and its diverse roles in cellular stress. Int. J. Biochem. Cell Biol. 2021, 133, 105934. [Google Scholar] [CrossRef] [PubMed]
  84. Pfaffenbach, K.T.; Lee, A.S. The critical role of GRP78 in physiologic and pathologic stress. Curr. Opin. Cell Biol. 2011, 23, 150–156. [Google Scholar] [CrossRef] [Green Version]
  85. Adachi, Y.; Yamamoto, K.; Okada, T.; Yoshida, H.; Harada, A.; Mori, K. ATF6 Is a Transcription Factor Specializing in the Regulation of Quality Control Proteins in the Endoplasmic Reticulum. Cell Struct. Funct. 2008, 33, 75–89. [Google Scholar] [CrossRef] [Green Version]
  86. Chen, Y.; Brandizzi, F. IRE1: ER stress sensor and cell fate executor. Trends Cell Biol. 2013, 23, 547–555. [Google Scholar] [CrossRef] [Green Version]
  87. Garcia de la Cadena, S.; Massieu, L. Caspases and their role in inflammation and ischemic neuronal death. Focus on caspase-12. Apoptosis 2016, 21, 763–777. [Google Scholar] [CrossRef]
  88. Shi, X.Y.; Li, J.T. Endoplasmic reticulum stress in regulation of hepatic fibrosis. Zhonghua Gan Zang Bing Za Zhi 2018, 26, 865–868. [Google Scholar]
  89. Zhang, P.; McGrath, B.; Li, S.A.; Frank, A.; Zambito, F.; Reinert, J.; Gannon, M.; Ma, K.; McNaughton, K.; Cavener, R.D. The PERK eukaryotic initiation factor 2 alpha kinase is required for the development of the skeletal system, postnatal growth, and the function and viability of the pancreas. Mol. Cell Biol. 2002, 22, 3864–3874. [Google Scholar] [CrossRef] [Green Version]
  90. Brewer, J.W.; Diehl, J.A. PERK mediates cell-cycle exit during the mammalian unfolded protein response. Proc. Natl. Acad. Sci. USA 2000, 97, 12625–12630. [Google Scholar] [CrossRef] [Green Version]
  91. You, K.; Wang, L.; Chou, C.-H.; Liu, K.; Nakata, T.; Jaiswal, A.; Yao, J.; Lefkovith, A.; Omar, A.; Perrigoue, J.G.; et al. QRICH1 dictates the outcome of ER stress through transcriptional control of proteostasis. Science 2021, 371, eabb6896. [Google Scholar] [CrossRef] [PubMed]
  92. Hetz, C.; Papa, F.R. The Unfolded Protein Response and Cell Fate Control. Mol. Cell 2018, 69, 169–181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Yoneda, T.; Imaizumi, K.; Oono, K.; Yui, D.; Gomi, F.; Katayama, T.; Tohyama, M. Activation of Caspase-12, an Endoplastic Reticulum (ER) Resident Caspase, through Tumor Necrosis Factor Receptor-associated Factor 2-dependent Mechanism in Response to the ER Stress. J. Biol. Chem. 2001, 276, 13935–13940. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Melber, A.; Haynes, C.M. UPR(mt) regulation and output: A stress response mediated by mitochondrial-nuclear communication. Cell Res. 2018, 28, 281–295. [Google Scholar] [CrossRef]
  95. Fessler, E.; Eckl, E.-M.; Schmitt, S.; Mancilla, I.A.; Meyer-Bender, M.F.; Hanf, M.; Philippou-Massier, J.; Krebs, S.; Zischka, H.; Jae, L.T. A pathway coordinated by DELE1 relays mitochondrial stress to the cytosol. Nature 2020, 579, 433–437. [Google Scholar] [CrossRef]
  96. Guo, X.; Aviles, G.; Liu, Y.; Tian, R.; Unger, B.A.; Lin, Y.-H.T.; Wiita, A.P.; Xu, K.; Correia, M.A.; Kampmann, M. Mitochondrial stress is relayed to the cytosol by an OMA1–DELE1–HRI pathway. Nature 2020, 579, 427–432. [Google Scholar] [CrossRef]
  97. Tian, X.; Zhang, S.; Zhou, L.; Seyhan, A.A.; Borrero, L.H.; Zhang, Y.; El-Deiry, W.S. Targeting the Integrated Stress Response in Cancer Therapy. Front. Pharmacol. 2021, 12, 747837. [Google Scholar] [CrossRef]
  98. Abdel-Nour, M.; Carneiro, L.A.M.; Downey, J.; Tsalikis, J.; Outlioua, A.; Prescott, D.; Da Costa, L.S.; Hovingh, E.S.; Farahvash, A.; Gaudet, R.G.; et al. The heme-regulated inhibitor is a cytosolic sensor of protein misfolding that controls innate immune signaling. Science 2019, 365, eaaw4144. [Google Scholar] [CrossRef]
  99. Area-Gomez, E.; Schon, E.A. Mitochondria-associated ER membranes and Alzheimer disease. Curr. Opin. Genet. Dev. 2016, 38, 90–96. [Google Scholar] [CrossRef] [Green Version]
  100. Aufschnaiter, A.; Kohler, V.; Diessl, J.; Peselj, C.; Carmona-Gutierrez, D.; Keller, W.; Büttner, S. Mitochondrial lipids in neurodegeneration. Cell Tissue Res. 2016, 367, 125–140. [Google Scholar] [CrossRef] [Green Version]
  101. van Vliet, A.R.; Giordano, F.; Gerlo, S.; Segura, I.; Van Eygen, S.; Molenberghs, G.; Rocha, S.; Houcine, A.; Derua, R.; Verfaillie, T.; et al. The ER Stress Sensor PERK Coordinates ER-Plasma Membrane Contact Site Formation through Interaction with Filamin-A and F-Actin Remodeling. Mol. Cell 2017, 65, 885–899.e6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Carreras-Sureda, A.; Jaña, F.; Urra, H.; Durand, S.; Mortenson, D.E.; Sagredo, A.; Bustos, G.; Hazari, Y.; Ramos-Fernández, E.; Sassano, M.L.; et al. Non-canonical function of IRE1α determines mitochondria-associated endoplasmic reticulum composition to control calcium transfer and bioenergetics. Nat. Cell Biol. 2019, 21, 755–767. [Google Scholar] [CrossRef] [PubMed]
  103. Bernard-Marissal, N.; Médard, J.-J.; Azzedine, H.; Chrast, R. Dysfunction in endoplasmic reticulum-mitochondria crosstalk underlies SIGMAR1 loss of function mediated motor neuron degeneration. Brain 2015, 138 Pt 4, 875–890. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Szymański, J.; Janikiewicz, J.; Michalska, B.; Patalas-Krawczyk, P.; Perrone, M.; Ziółkowski, W.; Duszyński, J.; Pinton, P.; Dobrzyń, A.; Więckowski, M.R. Interaction of Mitochondria with the Endoplasmic Reticulum and Plasma Membrane in Calcium Homeostasis, Lipid Trafficking and Mitochondrial Structure. Int. J. Mol. Sci. 2017, 18, 1576. [Google Scholar] [CrossRef] [Green Version]
  105. Ali, E.S.; Rychkov, G.; Barritt, G.J. Deranged hepatocyte intracellular Ca2+ homeostasis and the progression of non-alcoholic fatty liver disease to hepatocellular carcinoma. Cell Calcium 2019, 82, 102057. [Google Scholar] [CrossRef]
  106. Ali, E.S.; Rychkov, G.Y.; Barritt, G.J. Targeting Ca(2+) Signaling in the Initiation, Promotion and Progression of Hepatocellular Carcinoma. Cancers 2020, 12, 2755. [Google Scholar] [CrossRef]
  107. Hofmann, J.; Otarashvili, G.; Meszaros, A.; Ebner, S.; Weissenbacher, A.; Cardini, B.; Oberhuber, R.; Resch, T.; Öfner, D.; Schneeberger, S.; et al. Restoring Mitochondrial Function While Avoiding Redox Stress: The Key to Preventing Ischemia/Reperfusion Injury in Machine Perfused Liver Grafts? Int. J. Mol. Sci. 2020, 21, 3132. [Google Scholar] [CrossRef]
  108. Miller, B.A. TRPM2 in Cancer. Cell Calcium 2019, 80, 8–17. [Google Scholar] [CrossRef] [PubMed]
  109. Chen, X.; Zhang, L.; Zheng, L.; Tuo, B. Role of Ca2+ channels in non-alcoholic fatty liver disease and their implications for therapeutic strategies (Review). Int. J. Mol. Med. 2022, 50, 113. [Google Scholar] [CrossRef]
  110. Hudson, D.A.; Gannon, S.A.; Thorpe, C. Oxidative protein folding: From thiol–disulfide exchange reactions to the redox poise of the endoplasmic reticulum. Free. Radic. Biol. Med. 2015, 80, 171–182. [Google Scholar] [CrossRef] [Green Version]
  111. Li, Q.; Chen, F.; Wang, F. The immunological mechanisms and therapeutic potential in drug-induced liver injury: Lessons learned from acetaminophen hepatotoxicity. Cell Biosci. 2022, 12, 187. [Google Scholar] [CrossRef] [PubMed]
  112. Tasnim, F.; Huang, X.; Lee, C.Z.W.; Ginhoux, F.; Yu, H. Recent Advances in Models of Immune-Mediated Drug-Induced Liver Injury. Front. Toxicol. 2021, 3, 605392. [Google Scholar] [CrossRef]
  113. Shoda, L.K.; Battista, C.; Siler, S.Q.; Pisetsky, D.S.; Watkins, P.B.; Howell, B.A. Mechanistic Modelling of Drug-Induced Liver Injury: Investigating the Role of Innate Immune Responses. Gene Regul. Syst. Biol. 2017, 11, 1177625017696074. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Gerussi, A.; Natalini, A.; Antonangeli, F.; Mancuso, C.; Agostinetto, E.; Barisani, D.; Di Rosa, F.; Andrade, R.; Invernizzi, P. Immune-Mediated Drug-Induced Liver Injury: Immunogenetics and Experimental Models. Int. J. Mol. Sci. 2021, 22, 4557. [Google Scholar] [CrossRef] [PubMed]
  115. Verginadis, I.I.; Avgousti, H.; Monslow, J.; Skoufos, G.; Chinga, F.; Kim, K.; Leli, N.M.; Karagounis, I.V.; Bell, B.I.; Velalopoulou, A.; et al. A stromal Integrated Stress Response activates perivascular cancer-associated fibroblasts to drive angiogenesis and tumour progression. Nat. Cell Biol. 2022, 24, 940–953. [Google Scholar] [CrossRef]
  116. Gho, J.W.-M.; Ip, W.-K.; Chan, K.Y.-Y.; Law, P.T.-Y.; Lai, P.B.-S.; Wong, N. Re-Expression of Transcription Factor ATF5 in Hepatocellular Carcinoma Induces G2-M Arrest. Cancer Res. 2008, 68, 6743–6751. [Google Scholar] [CrossRef] [Green Version]
  117. Hu, H.; Tian, M.; Ding, C.; Yu, S. The C/EBP Homologous Protein (CHOP) Transcription Factor Functions in Endoplasmic Reticulum Stress-Induced Apoptosis and Microbial Infection. Front. Immunol. 2019, 9, 3083. [Google Scholar] [CrossRef] [Green Version]
  118. Zhang, T.; Huang, W.; Ma, Y. Down-regulation of TRPM2 attenuates hepatic ischemia/reperfusion injury through activation of autophagy and inhibition of NLRP3 inflammasome pathway. Int. Immunopharmacol. 2022, 104, 108443. [Google Scholar] [CrossRef]
  119. Kheradpezhouh, E.; Ma, L.; Morphett, A.; Barritt, G.J.; Rychkov, G.Y. TRPM2 channels mediate acetaminophen-induced liver damage. Proc. Natl. Acad. Sci. USA 2014, 111, 3176–3181. [Google Scholar] [CrossRef] [Green Version]
  120. Ali, E.S.; Rychkov, G.Y.; Barritt, G.J. TRPM2 Non-Selective Cation Channels in Liver Injury Mediated by Reactive Oxygen Species. Antioxidants 2021, 10, 1243. [Google Scholar] [CrossRef]
  121. Zhang, H.; Yu, P.; Lin, H.; Jin, Z.; Zhao, S.; Zhang, Y.; Xu, Q.; Liu, Z.; Yang, W.; Zhang, L. The Discovery of Novel ACA Derivatives as Specific TRPM2 Inhibitors that Reduce Ischemic Injury Both In Vitro and In Vivo. J. Med. Chem. 2021, 64, 3976–3996. [Google Scholar] [CrossRef] [PubMed]
  122. Bennett, B.L.; Sasaki, D.T.; Murray, B.W.; O’Leary, E.C.; Sakata, S.T.; Xu, W.; Leisten, J.C.; Motiwala, A.; Pierce, S.; Satoh, Y.; et al. SP600125, an anthrapyrazolone inhibitor of Jun N-terminal kinase. Proc. Natl. Acad. Sci. USA 2001, 98, 13681–13686. [Google Scholar] [CrossRef] [Green Version]
  123. Saito, C.; Lemasters, J.J.; Jaeschke, H. c-Jun N-terminal kinase modulates oxidant stress and peroxynitrite formation independent of inducible nitric oxide synthase in acetaminophen hepatotoxicity. Toxicol. Appl. Pharmacol. 2010, 246, 8–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Ren, Z.; Chen, S.; Qing, T.; Xuan, J.; Couch, L.; Yu, D.; Ning, B.; Shi, L.; Guo, L. Endoplasmic reticulum stress and MAPK signaling pathway activation underlie leflunomide-induced toxicity in HepG2 Cells. Toxicology 2017, 392, 11–21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Jang, S.; Yu, L.-R.; Abdelmegeed, M.A.; Gao, Y.; Banerjee, A.; Song, B.-J. Critical role of c-jun N-terminal protein kinase in promoting mitochondrial dysfunction and acute liver injury. Redox Biol. 2015, 6, 552–564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Loomba, R.; Lawitz, E.; Mantry, P.S.; Jayakumar, S.; Caldwell, S.H.; Arnold, H.; Diehl, A.M.; Djedjos, C.S.; Han, L.; Myers, R.P.; et al. The ASK1 inhibitor selonsertib in patients with nonalcoholic steatohepatitis: A randomized, phase 2 trial. Hepatology 2018, 67, 549–559. [Google Scholar] [CrossRef] [Green Version]
  127. Ibrahim, S.H.; Gores, G.J.; Hirsova, P.; Kirby, M.; Miles, L.; Jaeschke, A.; Kohli, R. Mixed lineage kinase 3 deficient mice are protected against the high fat high carbohydrate diet-induced steatohepatitis. Liver Int. 2014, 34, 427–437. [Google Scholar] [CrossRef] [Green Version]
  128. Min, R.W.M.; Aung, F.W.M.; Liu, B.; Arya, A.; Win, S. Mechanism and Therapeutic Targets of c-Jun-N-Terminal Kinases Activation in Nonalcoholic Fatty Liver Disease. Biomedicines 2022, 10, 2035. [Google Scholar] [CrossRef]
  129. Umbaugh, D.S.; Soder, R.P.; Nguyen, N.T.; Adelusi, O.; Robarts, D.R.; Woolbright, B.; Duan, L.; Abhyankar, S.; Dawn, B.; Apte, U.; et al. Human Wharton’s Jelly-derived mesenchymal stem cells prevent acetaminophen-induced liver injury in a mouse model unlike human dermal fibroblasts. Arch. Toxicol. 2022, 96, 3315–3329. [Google Scholar] [CrossRef]
  130. Luedde, T.; Schwabe, R.F. NF-κB in the liver--linking injury, fibrosis and hepatocellular carcinoma. Nat. Rev. Gastroenterol. Hepatol. 2011, 8, 108–118. [Google Scholar] [CrossRef] [Green Version]
  131. Huang, L.S.; Sudhadevi, T.; Fu, P.; Punathil-Kannan, P.-K.; Ebenezer, D.L.; Ramchandran, R.; Putherickal, V.; Cheresh, P.; Zhou, G.; Ha, A.W.; et al. Sphingosine Kinase 1/S1P Signaling Contributes to Pulmonary Fibrosis by Activating Hippo/YAP Pathway and Mitochondrial Reactive Oxygen Species in Lung Fibroblasts. Int. J. Mol. Sci. 2020, 21, 2064. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Bu, Y.; Wu, H.; Deng, R.; Wang, Y. Therapeutic Potential of SphK1 Inhibitors Based on Abnormal Expression of SphK1 in Inflammatory Immune Related-Diseases. Front. Pharmacol. 2021, 12, 733387. [Google Scholar] [CrossRef]
  133. Lima, S.; Takabe, K.; Newton, J.; Saurabh, K.; Young, M.M.; Leopoldino, A.M.; Hait, N.C.; Roberts, J.L.; Wang, H.-G.; Dent, P.; et al. TP53 is required for BECN1- and ATG5-dependent cell death induced by sphingosine kinase 1 inhibition. Autophagy 2018, 14, 942–957. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Pulkoski-Gross, M.J.; Uys, J.D.; Orr-Gandy, K.A.; Coant, N.; Bialkowska, A.B.; Szulc, Z.M.; Bai, A.; Bielawska, A.; Townsend, D.M.; Hannun, Y.A.; et al. Novel sphingosine kinase-1 inhibitor, LCL351, reduces immune responses in murine DSS-induced colitis. Prostaglandins Other Lipid Mediat. 2017, 130, 47–56. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Hengst, J.A.; Hegde, S.; Paulson, R.F.; Yun, J.K. Development of SKI-349, a dual-targeted inhibitor of sphingosine kinase and microtubule polymerization. Bioorganic Med. Chem. Lett. 2020, 30, 127453. [Google Scholar] [CrossRef]
  136. Alshaker, H.; Srivats, S.; Monteil, D.; Wang, Q.; Low, C.M.R.; Pchejetski, D. Field template-based design and biological evaluation of new sphingosine kinase 1 inhibitors. Breast Cancer Res. Treat. 2018, 172, 33–43. [Google Scholar] [CrossRef] [Green Version]
  137. Childress, E.S.; Kharel, Y.; Brown, A.M.; Bevan, D.R.; Lynch, K.R.; Santos, W.L. Transforming Sphingosine Kinase 1 Inhibitors into Dual and Sphingosine Kinase 2 Selective Inhibitors: Design, Synthesis, and in Vivo Activity. J. Med. Chem. 2017, 60, 3933–3957. [Google Scholar] [CrossRef]
  138. Vettorazzi, M.; Insuasty, D.; Lima, S.; Gutiérrez, L.; Nogueras, M.; Marchal, A.; Abonia, R.; Andújar, S.; Spiegel, S.; Cobo, J.; et al. Design of new quinolin-2-one-pyrimidine hybrids as sphingosine kinases inhibitors. Bioorganic Chem. 2020, 94, 103414. [Google Scholar] [CrossRef]
  139. Chen, H.-J.; Yang, H.-R.; Zhi, Y.; Yao, Q.-Q.; Liu, B. Evaluation of pyrrolidine-based analog of jaspine B as potential SphK1 inhibitors against rheumatoid arthritis. Bioorganic Med. Chem. Lett. 2021, 34, 127754. [Google Scholar] [CrossRef]
  140. Sun, D.; Wang, S. Sphingosine kinases are involved in the regulation of all-trans retinoic acid sensitivity of K562 chronic myeloid leukemia cells. Oncol. Lett. 2021, 22, 581. [Google Scholar] [CrossRef]
  141. Sah, R.K.; Pati, S.; Saini, M.; Singh, S. Erythrocyte sphingosine kinase regulates intraerythrocytic development of Plasmodium falciparum. Sci. Rep. 2021, 11, 1257. [Google Scholar] [CrossRef]
  142. Hafizi, R.; Imeri, F.; Wenger, R.H.; Huwiler, A. S1P Stimulates Erythropoietin Production in Mouse Renal Interstitial Fibroblasts by S1P1 and S1P3 Receptor Activation and HIF-2α Stabilization. Int. J. Mol. Sci. 2021, 22, 9467. [Google Scholar] [CrossRef] [PubMed]
  143. Hou, L.; Yang, L.; Chang, N.; Zhao, X.; Zhou, X.; Dong, C.; Liu, F.; Yang, L.; Li, L. Macrophage Sphingosine 1-Phosphate Receptor 2 Blockade Attenuates Liver Inflammation and Fibrogenesis Triggered by NLRP3 Inflammasome. Front. Immunol. 2020, 11, 1149. [Google Scholar] [CrossRef] [PubMed]
  144. Wang, L.-Y.; Sun, X.-J.; Wang, C.; Chen, S.-F.; Li, Z.-Y.; Chen, M.; Little, M.A.; Zhao, M.-H. Sphingosine-1-phosphate receptor modulator FTY720 attenuates experimental myeloperoxidase-ANCA vasculitis in a T cell-dependent manner. Clin. Sci. 2020, 134, 1475–1489. [Google Scholar] [CrossRef]
  145. Aye, I.L.; Waddell, B.J.; Mark, P.J.; Keelan, J.A. Oxysterols exert proinflammatory effects in placental trophoblasts via TLR4-dependent, cholesterol-sensitive activation of NF-κB. Mol. Hum. Reprod. 2012, 18, 341–353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Wu, Y.; Wang, Y.; Liu, B.; Cheng, Y.; Qian, H.; Yang, H.; Li, X.; Yang, G.; Zheng, X.; Shen, F. SN50 attenuates alveolar hypercoagulation and fibrinolysis inhibition in acute respiratory distress syndrome mice through inhibiting NF-κB p65 translocation. Respir. Res. 2020, 21, 130. [Google Scholar] [CrossRef]
  147. Yan, Y.; Qian, H.; Cao, Y.; Zhu, T. Nuclear factor-κB inhibitor Bay11-7082 inhibits gastric cancer cell proliferation by inhibiting Gli1 expression. Oncol. Lett. 2021, 21, 301. [Google Scholar] [CrossRef]
  148. Pao, H.-P.; Liao, W.-I.; Wu, S.-Y.; Hung, K.-Y.; Huang, K.-L.; Chu, S.-J. PG490-88, a derivative of triptolide, suppresses ischemia/reperfusion-induced lung damage by maintaining tight junction barriers and targeting multiple signaling pathways. Int. Immunopharmacol. 2019, 68, 17–29. [Google Scholar] [CrossRef]
  149. El-Hashim, A.Z.; Renno, W.M.; Abduo, H.T.; Jaffal, S.M.; Akhtar, S.; Benter, I.F. Effect of inhibition of the ubiquitin-proteasome-system and IκB kinase on airway inflammation and hyperresponsiveness in a murine model of asthma. Int. J. Immunopathol. Pharmacol. 2011, 24, 33–42. [Google Scholar] [CrossRef]
  150. Li, S.-Z.; Zhang, H.-H.; Zhang, J.-N.; Zhang, Z.-Y.; Zhang, X.-F.; Du, R.-L. ALLN hinders HCT116 tumor growth through Bax-dependent apoptosis. Biochem. Biophys. Res. Commun. 2013, 437, 325–330. [Google Scholar] [CrossRef]
  151. Kelly-Laubscher, R.; Somers, S.; Lacerda, L.; Lecour, S. Role of nuclear factor kappa-B in TNF-induced cytoprotection. Cardiovasc. J. Afr. 2022, 33, 1–7. [Google Scholar] [PubMed]
  152. Axten, J.M.; Medina, J.R.; Feng, Y.; Shu, A.; Romeril, S.P.; Grant, S.W.; Li, W.H.H.; Heerding, D.A.; Minthorn, E.; Mencken, T.; et al. Discovery of 7-Methyl-5-(1-{[3-(trifluoromethyl)phenyl]acetyl}-2,3-dihydro-1H-indol-5-yl)-7H-pyrrolo[2,3-d]pyrimidin-4-amine (GSK2606414), a Potent and Selective First-in-Class Inhibitor of Protein Kinase R (PKR)-like Endoplasmic Reticulum Kinase (PERK). J. Med. Chem. 2012, 55, 7193–7207. [Google Scholar] [CrossRef] [PubMed]
  153. Mahameed, M.; Wilhelm, T.; Darawshi, O.; Obiedat, A.; Tommy, W.-S.; Chintha, C.; Schubert, T.; Samali, A.; Chevet, E.; Eriksson, L.A.; et al. The unfolded protein response modulators GSK2606414 and KIRA6 are potent KIT inhibitors. Cell Death Dis. 2019, 10, 300. [Google Scholar] [CrossRef] [Green Version]
  154. Atkins, C.; Liu, Q.; Minthorn, E.; Zhang, S.-Y.; Figueroa, D.J.; Moss, K.; Stanley, T.B.; Sanders, B.; Goetz, A.; Gaul, N.; et al. Characterization of a Novel PERK Kinase Inhibitor with Antitumor and Antiangiogenic Activity. Cancer Res. 2013, 73, 1993–2002. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Rojas-Rivera, D.; Delvaeye, T.; Roelandt, R.; Nerinckx, W.; Augustyns, K.; Vandenabeele, P.; Bertrand, M.J.M. When PERK inhibitors turn out to be new potent RIPK1 inhibitors: Critical issues on the specificity and use of GSK2606414 and GSK2656157. Cell Death Differ. 2017, 24, 1100–1110. [Google Scholar] [CrossRef] [PubMed]
  156. Smith, A.L.; Andrews, K.L.; Beckmann, H.; Bellon, S.F.; Beltran, P.J.; Booker, S.; Chen, H.; Chung, Y.-A.; D’Angelo, N.D.; Dao, J.; et al. Discovery of 1H-Pyrazol-3(2H)-ones as Potent and Selective Inhibitors of Protein Kinase R-like Endoplasmic Reticulum Kinase (PERK). J. Med. Chem. 2015, 58, 1426–1441. [Google Scholar] [CrossRef] [PubMed]
  157. Wang, L.; Perera, B.G.K.; Hari, S.B.; Bhhatarai, B.; Backes, B.J.; Seeliger, M.A.; Schürer, S.C.; Oakes, S.A.; Papa, F.R.; Maly, D.J. Divergent allosteric control of the IRE1α endoribonuclease using kinase inhibitors. Nat. Chem. Biol. 2012, 8, 982–989. [Google Scholar] [CrossRef] [PubMed]
  158. Feldman, H.C.; Vidadala, V.N.; Potter, Z.E.; Papa, F.R.; Backes, B.J.; Maly, D.J. Development of a Chemical Toolset for Studying the Paralog-Specific Function of IRE1. ACS Chem. Biol. 2019, 14, 2595–2605. [Google Scholar] [CrossRef]
  159. Lebeau, P.; Byun, J.H.; Yousof, T.; Austin, R.C. Pharmacologic inhibition of S1P attenuates ATF6 expression, causes ER stress and contributes to apoptotic cell death. Toxicol. Appl. Pharmacol. 2018, 349, 1–7. [Google Scholar] [CrossRef]
  160. Torres, S.E.; Gallagher, C.M.; Plate, L.; Gupta, M.; Liem, C.R.; Guo, X.; Tian, R.; Stroud, R.M.; Kampmann, M.; Weissman, J.S.; et al. Ceapins block the unfolded protein response sensor ATF6alpha by inducing a neomorphic inter-organelle tether. eLife 2019, 8, e46595. [Google Scholar] [CrossRef]
  161. Ha, T.K.; Hansen, A.H.; Kildegaard, H.F.; Lee, G.M. BiP Inducer X: An ER Stress Inhibitor for Enhancing Recombinant Antibody Production in CHO Cell Culture. Biotechnol. J. 2019, 14, e1900130. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Mitochondrial stress and ERS in APAP and INH-induced liver injury. (APAP: Acetaminophen; NAPQI: N-acetyl-p-benzoquinone imine; ROS: Radical Oxygen Species; JNK: c-jun N-terminal kinase; MPT: Mitochondrial phosphate transporter; Bax: BCL2-Associated X; INH: Isoniazid; mit stress: Mitochondrial stress; mt DNA: Mitochondrial Deoxyribonucleic acid; GSH: Glutathione; ERS: Endoplasmic reticulum stress; TCA: Tricarboxylic acid).
Figure 1. Mitochondrial stress and ERS in APAP and INH-induced liver injury. (APAP: Acetaminophen; NAPQI: N-acetyl-p-benzoquinone imine; ROS: Radical Oxygen Species; JNK: c-jun N-terminal kinase; MPT: Mitochondrial phosphate transporter; Bax: BCL2-Associated X; INH: Isoniazid; mit stress: Mitochondrial stress; mt DNA: Mitochondrial Deoxyribonucleic acid; GSH: Glutathione; ERS: Endoplasmic reticulum stress; TCA: Tricarboxylic acid).
Molecules 28 03160 g001
Figure 2. SPHK1 pathway in liver injury. (SPHK1: sphingosine kinase-1; S1P: Sphingosine-1-phosphate; PERK: PKR-like ER kinase; IRE1α; Inositol-requiring enzyme-1α; CHOP: C/EBP-homologous Protein; ATF4: Activating Transcription Factor 4; ATF6: Activating Transcription Factor 6; ASK1: Apoptosis signal-regulating kinase 1; GSK3β: Glycogen Synthetase Kinase 3β; TRAF2: TNF receptor-associated factor 2; NF-κB: Nuclear Factor Kappa B).
Figure 2. SPHK1 pathway in liver injury. (SPHK1: sphingosine kinase-1; S1P: Sphingosine-1-phosphate; PERK: PKR-like ER kinase; IRE1α; Inositol-requiring enzyme-1α; CHOP: C/EBP-homologous Protein; ATF4: Activating Transcription Factor 4; ATF6: Activating Transcription Factor 6; ASK1: Apoptosis signal-regulating kinase 1; GSK3β: Glycogen Synthetase Kinase 3β; TRAF2: TNF receptor-associated factor 2; NF-κB: Nuclear Factor Kappa B).
Molecules 28 03160 g002
Figure 3. ATF6 pathway. (S2P: Sphingosine-1-phosphate; ERSE: ERS-response element; BIP: binding protein; XBP1: X-box binding protein 1).
Figure 3. ATF6 pathway. (S2P: Sphingosine-1-phosphate; ERSE: ERS-response element; BIP: binding protein; XBP1: X-box binding protein 1).
Molecules 28 03160 g003
Figure 4. IRE1 pathway. (S2P: Sphingosine-1-phosphate; ERSE: ERS-response element; BIP: binding protein; XBP1: X-box binding protein 1).
Figure 4. IRE1 pathway. (S2P: Sphingosine-1-phosphate; ERSE: ERS-response element; BIP: binding protein; XBP1: X-box binding protein 1).
Molecules 28 03160 g004
Figure 5. PERK pathway.
Figure 5. PERK pathway.
Molecules 28 03160 g005
Figure 6. Caspase-12 pathway.
Figure 6. Caspase-12 pathway.
Molecules 28 03160 g006
Figure 7. Mitochondrial stress pathways.
Figure 7. Mitochondrial stress pathways.
Molecules 28 03160 g007
Figure 8. Communication between the ERS and mitochondria stress (LONP1: Lon peptidase1; VDAC1: Voltage dependent anion channel protein 1; IP3R: Inositol 1,4,5 triphosphate receptor; MFN1: Mitofusin 1; MFN2: Mitofusin 2; Sig-1R: Sigma 1 receptor; DELE1: DAP3 binding cell death enhancer 1).
Figure 8. Communication between the ERS and mitochondria stress (LONP1: Lon peptidase1; VDAC1: Voltage dependent anion channel protein 1; IP3R: Inositol 1,4,5 triphosphate receptor; MFN1: Mitofusin 1; MFN2: Mitofusin 2; Sig-1R: Sigma 1 receptor; DELE1: DAP3 binding cell death enhancer 1).
Molecules 28 03160 g008
Table 1. Common drugs causing liver injury.
Table 1. Common drugs causing liver injury.
Drug ClassCompoundDrug TargetDamage TypeReferences
AntibioticsClindamycinHepatocyte; Bile duct epithelialHepatitis; Hepatic failure[23]
OxytetracyclineHepatocyte; Vascular injuryFatty degeneration[24]
CefazolinHepatocyteHepatitis[25,26]
NitrofurantoinHepatocyteHepatitis; Jaundice; Necrosis; Fibrosis[25]
Non-steroidal anti-inflammatory drugsAcetaminophenHepatocyteHepatonecrosis[4,5]
IbuprofenHepatocyte; Bile duct epithelial; Vascular injuryHepatocyte damage; Hepatic failure; Acute hepatitis[27,28]
AspirinHepatocyte; Bile duct epithelial; Vascular injuryHepatic enzymes elevations; Jaundice; Acute hepatitis[28]
DiclofenacHepatocyteHepatitis; Jaundice; Hepatauxe[29]
AntipsychoticsRisperidoneHepatocyte; Bile duct epithelialHepatitis[30,31]
ClozapineHepatocyte; Bile duct epithelialFulminant liver failure[32]
AntidepressantAmitriptylineHepatocyte; Bile duct epithelialHepatitis; Fulminant liver failure or death[31]
Trazodone
Agomelatine
Bupropion
Iproniazid
Duloxetine
Imipramine
Nefazodone
Tianeptine
Phenelzine
Lipid-lowering drugsAtorvastatinHepatocyte; Bile duct epithelial; Vascular injuryHepatitis; Jaundice[33]
ChemotherapeuticsFluorouracilHepatocyteHepatitis; Steatosis[34,35]
IrinotecanHepatocyteHepatitis; Steatosis[34]
OxaliplatinHepatocyte; Vascular injuryHepatic sinusoidal obstruction syndrome[36]
Antituberculosis drugsIsoniazidHepatocyte; Bile duct epithelialHepatitis; Necrosis[6,7,8]
RifampicinHepatocyte; Bile duct epithelialHepatitis; Necrosis[6,7,37]
Pyrazinamide
(PZA)
Hepatocyte; Vascular injuryHepatitis; Necrosis[38,39]
Antiepileptic drugsLamotrigineHepatocyteHepatic enzymes elevations; Death; Necrosis[40]
Valproic acid (VPA)HepatocyteHepatitis;[41,42]
Traditional Chinese medicine (TCM)HeShouWuHepatocyteHepatitis; Jaundice; Necrosis[43]
JuSanQiHepatocyte; Vascular injuryHepatitis; Hepatic sinusoidal obstruction syndrome[3]
Dioscorea bulbifera L.Hepatocyte; Bile duct epithelial; Vascular injuryHepatitis[44]
CangErZiHepatocyte; Vascular injuryHepatitis[45]
Oral anticoagulantsRivaroxabanHepatocyte; Bile duct epithelial; Vascular injuryHepatitis; Hepatic enzymes elevations; Jaundice[46,47]
ApixabanHepatocyte; Bile duct epithelial;Hepatitis[48,49]
Antiandrogen drugFlutamideHepatocyteHepatitis; Jaundice; ALF[50,51]
AntimicrobialKetoconazoleHepatocyteLiver cirrhosis; Jaundice[52]
Hypoglycemic agentsTroglitazone *HepatocyteALF[53]
Acid-inhibitory drugsEsomeprazoleHepatocyteHepatic enzymes elevations[54]
CimetidineHepatocyteJaundice; ALF[55]
* Troglitazone has been discontinued due to its severe hepatotoxicity. When applying such drugs, liver function should be closely monitored.
Table 2. Pathway inhibitors in DILI.
Table 2. Pathway inhibitors in DILI.
PathwayClassCompoundCommentsTargetReferences
SPHK1-STARD pathwaySPHK1 inhibitorsPF543Sphingosine kinase inhibitorSPHK1[14,131]
SK1-IEnhances autophagy and cancer cell deathSPHK1[132,133]
LCL351Reduces the expression of pro-inflammatory cytokineSPHK1[134]
SKI-178/349High selectivity and low toxicityATP- binding site[135]
SK-FWithout significant systemic toxicitySPHK1[136]
SLC4011540High cell permeabilitySPHK1[137]
11bSelective inhibition of SPHK1SPHK1[138]
CHJ01Anti-inflammatorySPHK1[139]
SK1-IIUp-regulates Cer levelSPHK1[140]
DMSCauses severe hemolysis in miceSPHK1[141]
S1P inhibitorsNIBR0213Reduces peripheral blood lymphocyte countsS1P1[142]
JTE013Inhibits inflammasome priming and inflammatory cytokineS1P2[143]
FTY720Reduces and inhibits T cellsS1P3[144]
TRPM2 pathwayTRPM2 inhibitorsCurcuminPrevents ROS-induced liver injuryTRPM2[120]
Compound A23Highly active and selectiveTRPM2[121]
JNK pathwayJNK inhibitorSP600125Inhibits mitochondrial stressJNK1/2/3[123]
LEFInhibits JNK1/2 phosphorylationJNK1/2[124]
BI-78D3Binds D-domain of JIP1D-domain of JIP1[125]
ASK1 inhibitorGS-444217Noneffective in late-phase clinical trialsASK1[126]
NF-κB pathwayIKK inhibitorsParthenolideSpecific IkB inhibitorIkB[145]
NF-κB inhibitorsSN50NF-κB inhibitory peptides for permeable cellsNF-κB[146]
BAY11-7082Prevents IkBα phosphorylationNF-κB[147]
Immunosuppressive agentsPG490Target gene transcription inhibitorP65[148]
Protease inhibitorsMG-132Inhibits phosphorylation of IkBαIkBα[149]
AntioxidantsALLNPrevents IkBα degradationIkBα[150]
PDTCPrevents IkB releaseIkB[151]
UPR pathwayPERK inhibitorsGSK2606414RIPK inhibitor; KIT inhibitorEIF2AK3[152,153]
GSK2656157RIPK inhibitorATP- binding site[154,155]
PERK-IN-2Low renal clearanceATP- binding site[152]
PERK-IN-3High renal clearanceATP- binding site[152]
AMG PERK 44GCN2 inhibitorPERK[156]
IRE1 inhibitorsCompound 3Compound 3 disfavors IRE1 oligomerizationSTING[157,158]
Compound 15EGFR inhibitorEGFR[157,158]
ATF6 inhibitorsPF-429242Inhibits SREBP signalingS1P[159]
CeapinDoes not affect ABCD3 functionTethers ATF6 cytosolic
domain and ABCD3
[160]
Bip inducerBIXERS inhibitorBip[161]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Pu, S.; Pan, Y.; Zhang, Q.; You, T.; Yue, T.; Zhang, Y.; Wang, M. Endoplasmic Reticulum Stress and Mitochondrial Stress in Drug-Induced Liver Injury. Molecules 2023, 28, 3160. https://doi.org/10.3390/molecules28073160

AMA Style

Pu S, Pan Y, Zhang Q, You T, Yue T, Zhang Y, Wang M. Endoplasmic Reticulum Stress and Mitochondrial Stress in Drug-Induced Liver Injury. Molecules. 2023; 28(7):3160. https://doi.org/10.3390/molecules28073160

Chicago/Turabian Style

Pu, Sisi, Yangyang Pan, Qian Zhang, Ting You, Tao Yue, Yuxing Zhang, and Meng Wang. 2023. "Endoplasmic Reticulum Stress and Mitochondrial Stress in Drug-Induced Liver Injury" Molecules 28, no. 7: 3160. https://doi.org/10.3390/molecules28073160

Article Metrics

Back to TopTop