Next Article in Journal
Peroxisomal Localization of a Truncated HMG-CoA Reductase under Low Cholesterol Conditions
Previous Article in Journal
Current Status and Future Prospects of Stem Cell Therapy for Infertile Patients with Premature Ovarian Insufficiency
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of the Dysregulated JNK Signaling Pathway in the Pathogenesis of Human Diseases and Its Potential Therapeutic Strategies: A Comprehensive Review

1
Department of Ultrasound, Hospital of Chengdu University of Traditional Chinese Medicine, Chengdu 610075, China
2
Department of Endocrinology, Hospital of Chengdu University of Traditional Chinese Medicine, Chengdu 610075, China
3
Cancer Center and State Key Laboratory of Biotherapy, Department of Biotherapy, West China Hospital, Sichuan University, Chengdu 610041, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Biomolecules 2024, 14(2), 243; https://doi.org/10.3390/biom14020243
Submission received: 6 December 2023 / Revised: 12 February 2024 / Accepted: 15 February 2024 / Published: 19 February 2024

Abstract

:
JNK is named after c-Jun N-terminal kinase, as it is responsible for phosphorylating c-Jun. As a member of the mitogen-activated protein kinase (MAPK) family, JNK is also known as stress-activated kinase (SAPK) because it can be activated by extracellular stresses including growth factor, UV irradiation, and virus infection. Functionally, JNK regulates various cell behaviors such as cell differentiation, proliferation, survival, and metabolic reprogramming. Dysregulated JNK signaling contributes to several types of human diseases. Although the role of the JNK pathway in a single disease has been summarized in several previous publications, a comprehensive review of its role in multiple kinds of human diseases is missing. In this review, we begin by introducing the landmark discoveries, structures, tissue expression, and activation mechanisms of the JNK pathway. Next, we come to the focus of this work: a comprehensive summary of the role of the deregulated JNK pathway in multiple kinds of diseases. Beyond that, we also discuss the current strategies for targeting the JNK pathway for therapeutic intervention and summarize the application of JNK inhibitors as well as several challenges now faced. We expect that this review can provide a more comprehensive insight into the critical role of the JNK pathway in the pathogenesis of human diseases and hope that it also provides important clues for ameliorating disease conditions.

1. Introduction

Cells are able to communicate with each other and adjust their behaviors to adapt to environmental changes. Extracellular signaling is transduced to intracellular organelles through multiple pathways [1]. Among these signaling networks, MAPK cascades are involved in transmitting extracellular signals to intracellular targets [2,3,4]. The MAPK family includes p38, ERK, and JNK. In general, the MAPK signaling pathway mainly consists of MAP3K, MAP2K, and MAPK [5]. Functionally, MAPK signaling regulates several cellular behaviors such as cell proliferation, survival, and death [5]. However, when it is deregulated, MAPK signaling can aggravate the progression of some diseases, such as metabolic disorders, neurological diseases, and various types of cancer [5].
JNK (also known as SAPK) is named after c-Jun N-terminal kinase, as it phosphorylates the N-terminal serine residues and activates c-Jun transcription factor [6,7,8]. The JNK family includes JNK1, JNK2, and JNK3. These three isoforms are encoded by the jnk1, jnk2, and jnk3 genes, respectively [9]. Like other MAPK family members, JNK also regulates a variety of cell processes such as apoptosis, metabolic reprogramming, and inflammation [9,10,11,12]. JNK activity is strictly regulated in most cells. Therefore, deregulated JNK signaling may cause different, even antagonistic, functions of JNK signaling; for example, impermanent JNK signaling contributes to cell survival, while durative JNK activation leads to cell death [1,13].
When it is dysregulated, JNK signaling is harmful to cells and ultimately to the body. JNK over-activation is involved in the development of a variety of diseases. Although the role of the JNK signaling pathway in a single disease such as diabetes has been summarized in several previous publications, a comprehensive review of the JNK signaling pathway and its roles in multiple kinds of diseases is missing. Therefore, in this work, we attempt to comprehensively summarize the recent advances in the JNK research area. To this end, we first discuss the landmark discoveries, structures, tissue expressions of JNK isoforms, and JNK’s activation mechanism. Next, we focus on the effects of the dysregulated JNK pathway on several types of human diseases, including metabolic disorders, neurodegenerative diseases, chronic inflammation and autoimmune diseases, cancers, infectious diseases, and other diseases. Lastly, we summarize the current therapeutic strategies for targeting JNK in treating these diseases. We expect that this review can provide a more comprehensive insight into the critical role of the JNK pathway in multiple kinds of human diseases and hope that it also provides important clues for ameliorating disease conditions.

2. Overview of the JNK Signaling Pathway

2.1. Landmark Discoveries of the JNK Signaling Pathway

JNKs were initially discovered in the early 1990s [14,15,16,17]. Subsequently, the three mammalian JNK isoforms were cloned in succession [17,18,19], and the functional significance of JNKs, including embryonic development and neurodegeneration, was further explored. Then, researchers have taken great efforts to explore the characteristics of JNKs in the past two decades. These efforts included the structural determination of JNK1, JNK2, and JNK3 [20,21,22], the finding of JNK scaffolds [23,24], and the development of small-molecule inhibitors. Landmark discoveries of the JNK pathway were summarized in Figure 1. Notably, both the structural determination of JNKs in complex with their ligands and the exploration of JNK isoform-specific inhibitors have become two important issues in recent years.

2.2. Alternative Splicing, Typical Structure of JNK Isoforms, and Their Expression Localizations

JNK is an evolutionarily conserved family of serine/threonine protein kinases. The sequences of the JNK isoform-coding genes are highly conserved through evolution [25]. Alternative gene splicing is now recognized as a mechanism to produce different splicing forms. Correspondingly, alternative splicing of the Jnk gene produces multiple isoforms of the JNK gene [19,26]. In general, four spliced forms (α1, α2, β1, and β2) are generated from the Jnk1 and Jnk2 genes, respectively, and eight from the Jnk3 gene [27,28,29].
As shown in Figure 2A, alternative splicing of JNK1-α1/α2/β1/β2 and JNK2-α1/α2/β1/β2 mainly occurs at alternative exon 6, resulting in long and short N- or C-termini, while the sequence differences of JNK3-α1 and JNK3-α2 are the lengths of the N- or C-termini. The JNK isoforms are typical serine/threonine protein kinases, containing all 11 subdomains as indicated (I–XI). Figure 2B shows the protein sequence of human JNK1, in which the protein kinase activation loop is located between domains VII and VIII. The kinase activation loop contains the conserved amino acid motifs DFG and APE. Within this loop is the threonine (T)-Pro (P)-tyrosine (Y) motif. Structurally, JNK1 contains the N-terminal lobe, the C-terminal lobe, and the activation loop. Catalytic sites are located in the activation loop (Figure 2C). Notably, there is one splicing site between subdomains IX and X of the JNK1 and JNK2 genes, and the resulting splice forms show altered substrate specificity (Figure 2D). The second alternative splicing occurs at the C-terminus of the protein, producing proteins that differ in length by 42 or 43 amino acids (i.e., the typical 46 and 55 kDa proteins). These alternative splicing isoforms cause some functional changes, such as changes in alternative substrate binding, changes in protein interaction, and changes in expression levels [30,31]. Thus, the isoforms of JNK have different substrate specificities and can phosphorylate different non-nuclear substrates or downstream nuclear transcription factors [26,27,32]. As for tissue localization, JNK1/2 expresses in any one cell and in various tissues throughout the body, whereas JNK3 mainly expresses in a few tissues, including the brain, testis, heart, and pancreas [18,27,33,34,35,36]. In terms of cellular localization, although JNK MAPKs are located in both the cytoplasm and nucleus of quiescent cells, activated JNK MAPKs can translocate to the nucleus [31]. Specifically, JNK1 and JNK2 mainly occur in the cytoplasm and nucleus [37,38]. JNK3 can occur in the cytoplasm, nucleus, membrane, and mitochondrion [39]. Due to the differences in cellular and tissue localization, different JNK isoforms may play different roles in the onset and progression of different tissue-related diseases.

2.3. Structures of JNK Isoforms

The molecular weights of the three JNK isoforms (JNK1, JNK2, and JNK3) are either 46 kDa or 55 kDa. There is a COOH-terminal extension in some JNK isoforms [40]. The crystal structures of the three JNK isoforms were determined, showing over 90% sequence homology [20,21,22]. Structurally, the JNK protein mainly includes three parts, i.e., the N-terminal lobe, the C-terminal lobe, and a flexible segment connecting the two lobes. The N-lobe contains multiple β strands, while the C-lobe is rich in α helices [27]. Furthermore, the former mainly contains several glutamate–aspartate (ED) domains, while the latter chiefly consists of the common docking (CD) domain [27,41]. The flexible peptide segment provides the kinase activity center containing a Thr-Pro-Tyr motif, which can be phosphorylated by MAP2K kinase. Additionally, JNK kinase includes a kinase activation site and a highly conserved ATP-binding site, which causes a lack of selectivity for small-molecule inhibitors [27]. Figure 2C shows the typical structure of the human JNK1 protein. Recently, the crystal structures of JNK1 bound to the MKK7 docking motif (D2 peptide) [42], JNK1 in complex with ATF2 [43], and AMP-bound human JNK2 [44], as well as JNK3 complexed with thiophene–pyrazolourea derivatives [45], were also reported and shown in Figure 3A, B, C, and D, respectively. These structural determinations will help explore novel JNK inhibitors.

2.4. General Activation Mechanism of the JNK Signaling Pathway

JNK belongs to the prodirected serine/threonine kinase family and can be activated by diverse external factors, including cytokines, growth factors, ROS, heat shock, shear stress, pathogens, and drugs [36,46]. Like other MAPK pathways, the JNK pathway is mainly composed of three components, i.e., MAP3Ks (e.g., ASK1), MAP2Ks (i.e., MKK4 and MKK7), and MAPK (JNK1/2/3). As shown in Figure 4, when a cell is stimulated by extracellular stresses, MAP3K (i.e., MEKK1) is first activated in the JNK pathway. Then, the activated MAP3K phosphorylates MKK4 and MKK7, two members of the MAP2K family, which finally phosphorylate and activate the MAPKs (JNK1, JNK2, and JNK3) [28]. Like other MAPK family members, JNK is activated when the Thr-Pro-Tyr tripeptide motif (T-X-Y motif) in the flexible segment is dually phosphorylated by MAP2K. For example, JNK1 is activated when dual phosphorylation occurs on Thr-183 and Tyr-185 [9,26,28]. Although the process of JNK activation is relatively clear, an important question whether the activation of JNKs is due to an increased level of kinase activity or due to an elevated level of JNK proteins need to be addressed. Thus, we decided to discuss this issue in the following several paragraphs.
Previous studies have shown that the catalytic activity of MAPK protein kinase is tightly regulated through the activation segment that is phosphorylated by other kinases to facilitate catalytic activity [47,48]. In general, a kinase will phosphorylate the activation segment of a downstream kinase, allowing the downstream kinase to further propagate a signal. Importantly, phosphorylation of the activation segment at the primary phosphorylation site stabilizes the kinase in a conformation suitable for substrate binding. Kinases may also have secondary phosphorylation sites in their activation segment, which may enhance their activity. Secondary phosphorylation may also aid in the recruitment of substrates by changing the conformation of a kinase to facilitate substrate binding. Once a kinase is stabilized and activated, the catalytic domain identifies a specific substrate and phosphorylates the substrate via its active site [47]. In other words, phosphorylation of the activation segment plays an important role in the regulation of kinase activity. For example, the Ste20-like kinase (SLK), a JNK upstream kinase, can phosphorylate JNK. Luhovy et al. found that, compared with WT-SLK (wild-type SLK), mutations in serine (Ser-189) and threonine (Thr-183) residues in the activation segment of SLK significantly reduce its kinase activity. That is, the T183A, S189A, and T183A/S189A mutants show reduced in vitro kinase activity of SLK. Wild-type SLK, but not mutants, increases activation-specific phosphorylation of JNK kinase [48]. These findings suggest that the phosphorylation of serine (Ser-189) and threonine (Thr-183) residues in the activation segment of SLK determines its kinase activity, which is responsible for phosphorylating its downstream JNK.
As discussed above, JNK belongs to the mitogen-activated protein kinase (MAPK) family. Similar to SLK, JNK activation requires the phosphorylation of both Thr and Tyr residues in its Thr-Pro-Tyr motif in the activation loop between VII and VIII of the kinase domain. For example, JNK1 activation requires the dual phosphorylation of Tyr-185 and Thr-183 in the Thr-Pro-Tyr motif, which is sequentially catalyzed by MKK4 and MKK7, respectively. Interestingly, MKK4 shows a striking preference for the tyrosine residue (Tyr-185), and MKK7 shows a significant preference for the threonine residue (Thr-183) [49,50,51]. From these findings, we can draw the conclusion that, similar to SLK, activation of JNKs also requires the phosphorylation of both Thr and Tyr residues in the activation loop.
Previous studies have shown that the phosphorylation of JNK at Thr-183 and Tyr-185 (p-Thr183/Tyr185) represents its activation form, which exerts the kinase activity of JNK and can phosphorylate its substrates, including c-Jun [52,53,54]. For example, Christopher et al. have recently reported that the phosphorylated JNK (p-Thr183/Tyr185) shows kinase activity on c-Jun and phosphorylates the c-Jun protein at four residues within its transactivation domain (TAD). Among these residues, Ser-63 and Ser-73 are phosphorylated by JNK more rapidly than Thr-91 and Thr-93. Interestingly, different c-Jun phosphorylation states exert different functions: Unphosphorylated c-Jun recruits the MBD3 repressor; Ser-63/73 doubly-phosphorylated c-Jun binds to the TCF4 co-activator, whereas the fully phosphorylated form disfavors TCF4 binding, attenuating JNK signaling. That is, c-Jun phosphorylation encodes multiple functional states that drive a complex signaling response from a single JNK input [54].
To the best of our knowledge, no report has shown that the activation of JNKs is directly related to an elevated level of JNK proteins. However, Bildik et al. found that siRNAs targeting JNK, which significantly downregulated JNK expression, led to a significant downregulation of p-c-Jun (Ser-63) and p-c-Jun (Ser-73) [53], implying that the expression level of JNK had an effect on the activation of JNK. Additionally, Yue et al. found that JNK was activated early during stress-induced apoptosis [55,56], and sustained JNK activation accelerated the apoptotic program. Importantly, when caspase-3 was activated during the apoptotic process, JNK was proteolyzed at Asp-385, increasing the release of cytochrome c and caspase-3 activity, thereby creating a positive feedback loop [56]. These studies indicate that the amount of JNK protein may affect its activation. Based on these findings, we think that the activation of JNKs is due to increased kinase activity; however, the expression level of JNK proteins may have an effect on the activation of JNKs.
In the upstream of the JNK pathway, MKK4/7 is regulated by its upstream MAP3Ks. Similarly, MAP3Ks are also activated upon the phosphorylation of specific serine or threonine residues [26]. The common MAP3K members mainly include MLK, ASK1, TAK1, and TPL2 [9]. Notably, scaffold proteins, for instance, JIP1-3, POSH, and IKAP [57,58], facilitate the sequential phosphorylation cascade, while dual specificity protein phosphatase (DUSP) deactivates this pathway through dephosphorylation [28,59]. Additionally, MEKK1 upstream kinases such as NIK, HPK, and GCK are also involved in regulating the activation of the JNK pathway through the phosphorylation of MEKK1. These findings suggest that DUSPs negatively, while HPK and GCK positively, regulate JNK signaling.
The mechanism by which external factors activate JNK is relatively clear. However, how pathogens cause the activation of JNK is not fully understood. Thus, we take some viruses as examples to illustrate which factors of pathogens are decisive for activating the JNK pathway. Previous studies have shown that JNK can be activated by some viruses, such as influenza A virus and HIV-1 [60,61,62]. Mechanistically, most of these viruses use their viral proteins to activate JNK signaling. For instance, the NS1 protein of influenza A virus can activate JNK. Moreover, the amino acid residue phenylalanine (F) at position 103 of NS1 is decisive for JNK activation [60,63]. Similarly, the Tat protein of HIV-1 (human immunodeficiency virus, type 1) activates JNK signaling through a Nox2-dependent manner [62]. SARS-CoV-2 virus also uses its spike protein to activate the JNK pathway through toll-like receptor signaling [64]. Based on these findings, we think that viral proteins are critical for the activation of JNKs.
Once activated, JNK then phosphorylates its downstream target proteins, which are involved in regulating different cellular responses. Up to now, JNK can phosphorylate almost 100 well-defined substrates [59]. These downstream targets mainly include scaffold proteins (e.g., JIP1), mitochondrial proteins, cytoskeletal proteins, transcription factors (e.g., ATF2), and transmembrane receptors [59,65]. Due to the involvement of these substrates in signaling transduction [59], the activated JNK can regulate a number of cell processes, such as cell differentiation, proliferation, and survival, which play different roles in many types of diseases, such as cancer, metabolic disorders, neurodegenerative diseases, chronic inflammation, autoimmune diseases, and infectious diseases. Notably, several JNK protein complexes are also involved in JNK signaling and some diseases. For instance, the complex composed of ACID, JIP1, and JNK1, the complex composed of JIP1 and JNK1/2/3, and the complex composed of Sab and JNK1/2/3 are also involved in JNK signaling transduction and are critical for neurodegenerative diseases, autoimmune inflammatory diseases/glomerulosclerosis, and non-alcoholic steatohepatitis, respectively. The TAT-JIP peptide and Notch 1-IC (Notch intracellular domain) can be used for ameliorating autoimmune inflammatory diseases and glomerulosclerosis, respectively, through a competitive inhibition of the interaction of JIP1 and JNK1/2/3.
As for the transcription factors phosphorylated by JNK, it is worth noting that c-Jun is an exclusive JNK MAPK substrate. However, JNK can also phosphorylate and activate other transcription factors. ATF2 is a JNK MAPK substrate that heterodimerizes with c-Jun and stimulates the expression of the c-Jun gene [31]. Therefore, through the activation of both c-Jun and ATF2, JNK can regulate the abundance and activity of c-Jun. Elk-1 is another direct JNK MAPK target, whose product forms the AP-1 heterodimer with c-Jun. In the cases of the transcription factors ATF2 and Elk-1, it should also be noted that other protein kinases can lead to their phosphorylation and activation. Specifically, ATF2 can also be phosphorylated by p38 MAPKs, whereas Elk-1 may also be phosphorylated by both ERK and p38 MAPKs [31]. Thus, there is the possibility of cross-talk between the MAPK pathways at the levels of their transcription factor substrates. These findings show that c-Jun may be the common target for JNKs and that JNK MAPKs can phosphorylate various targets to regulate their activity, thus eliciting important biological effects.
In addition, from the activation mechanism of the JNK signaling pathway described above, we can draw the conclusion that the commonalities of JNKs include, but are not limited to, the following elements: (1) different JNK isoforms have a similar structure (e.g., N-lobe, C-lobe, and activation loop); (2) the JNK isoforms all belong to the typical serine/threonine protein kinase, which contains all 11 subdomains; (3) all of the JNK isoforms can phosphorylate c-Jun; (4) they can mediate cell survival or cell death; and (5) JNK isoforms can be negatively regulated by DUSPs or PP2A and regulated by JIP.
Next, we would like to briefly summarize the regulation of JNK signaling. As we know, JNK belongs to the serine/threonine protein kinase family. As an enzyme, the kinase activity of JNK is strictly regulated in the cell. When the JNK signaling pathway is disturbed, i.e., the kinase activity affected by deviations in the common pathways of activation and inactivation within cells, the deregulated JNK signaling is harmful to cells. The JNK signaling pathway can be disturbed by external factors, such as cytokines, growth factors, ROS, pathogens, and drugs. The abnormal expressions of some genes in the JNK pathway or mutations can also disturb JNK signaling. For instance, the constitutively active mutant of MKK4 can cause sustained activation of JNK, while their dominant-negative mutants can inactivate JNK kinase. By contrast, DUPS overexpression in cells can lead to the inactivation of JNK. In addition, JIP-mimicry peptides can inactivate JNK by interfering with the interaction of JNK and JIP. For example, over-expressed JIP-1 can inhibit JNK MAPK signaling either by inhibiting JNK activity or by altering the subcellular localization of JNK [31]. Interestingly, no report, to the best of our knowledge, has shown that mutations of JNKs themselves lead to the activation of JNKs; however, the constitutively active mutations of some upstream kinases (e.g., TAK1 or MKK4/7) or other molecules (e.g., agouti-related peptide) can activate JNKs (AgRP) [66,67]. In addition, small-molecule inhibitors can inactivate JNK kinase by occupying the kinase catalytic center, inhibiting the kinase activity, or interfering with the JNK–JIP interaction. Due to the complexity of the regulating mechanism of JNK activation and the damage to cells caused by deregulated JNK signaling, deregulated JNK signaling may play a critical role in human diseases. In the following section, we will discuss the role of dysregulated JNK signaling in human diseases.

3. The Role of the Dysregulated JNK Pathway in Human Diseases

As discussed above, JNK activity is strictly regulated within cells. Once the JNK pathway is dysregulated, it is harmful to cells and ultimately aggravates disease conditions. Dysregulated JNK signaling contributes to a variety of diseases involving metabolic disorders, chronic inflammation, autoimmune diseases, neurodegeneration, cancer, infectious diseases, and other diseases such as ischemia/reperfusion injury, hearing loss/deafness, and kidney fibrosis. In this section, we mainly discuss the role of the dysregulated JNK signaling pathway in these diseases.

3.1. The Role of the Dysregulated JNK Pathway in Metabolic Disorders

Metabolic syndrome, a metabolic disorder-associated disease, mainly includes obesity, type 2 diabetes (T2D), cardiovascular disorders, and non-alcoholic steatohepatitis (NASH). The dysregulated JNK pathway plays a role in the onset and progression of metabolic syndromes.

3.1.1. The JNK Pathway in Obesity

It is well known that obesity is caused by excessive fat accumulation [68,69]. Obesity often leads to several health problems, such as insulin resistance and T2D [70], fatty liver disease, cardiovascular disease, and cancer [71]. Currently, obesity-driven diabetes is becoming a global metabolic disorder [26]. Although the mechanisms for obesity-triggering T2D are not fully understood, inflammatory kinases play a role in insulin resistance [68]. As an inflammatory kinase, JNK is activated by pro-inflammatory cytokines and mediates the transition from obesity to T2D [68].
Previous studies have shown that obesity is a chronic inflammatory disease [36,68]. The evidence is as follows: a large number of macrophages are observed in the visceral adipose tissue. Moreover, macrophages in adipose tissues also secrete pro-inflammatory cytokines, including TNFα and IL-1, which then activate JNK kinases [36]. In addition, JNKs are also stimulated by high-fat diets and are activated by TNFα and IL-1 [26,72,73]. Obesity is also associated with the increase in adipocyte mass, adipokine secretion, and free fatty acids (FFAs), all of which can activate JNK kinases [27]. For instance, FFAs cause JNK activation in cultured 3T3-L1 cells [74]. In adipose tissue, FFAs activate JNK by promoting the activation of the Src–MLK axis [75,76]. Additionally, the scaffold protein JIP1 has been shown to link the Src–MLK-JNK axis with the FFA–JNK signaling axis [75,76]. Importantly, JIP−/− and MLK−/− mice were protected from obesity, showing that the JIP1-mediated Src–MLK signaling axis is needed for obesity [75,77,78,79]. Overall, these findings suggest that the JNK pathway plays a critical role in obesity.

3.1.2. The JNK Pathway in Insulin Resistance and T2D

High-fat diets (HFDs) trigger JNK over-activation, which leads to insulin resistance [80]. The opinion that JNK functions in insulin resistance is supported by the following observations: (1) JNK over-activation is observed during obesity, and mice with a JNK1 deficiency display a significantly improved insulin sensitivity level [72,80]; (2) JNK1 knockout mice are protected from obesity [80,81]; (3) mice lacking JNK2 also display an improved insulin sensitive phenotype [81]; and (4) there are some mutations in the JIP1 coding gene (JNK’s scaffold) in patients with type II diabetes [82].
As for the action mechanism, Solinas et al. provided four distinct mechanisms to explain the involvement of JNK in T2D: (1) the phosphorylation of IRS1/2; (2) contribution to metabolic inflammation; (3) negative regulation of the TSH–thyroid hormone axis; and (4) inhibition of PPARa-FGF21 signaling [80]. These four mechanisms are summarized in Figure 5. The first mechanism is supported by the report that anti-TNFα treatment ameliorates insulin sensitivity [83] and also by the findings that TNFα-activated JNK directly phosphorylates IRS1/2, causing insulin resistance [80,84]. Specifically, Tanti et al. found that phosphorylation of IRS1 by JNK inhibits its interaction with insulin receptor (IR), causes defective IRS1 tyrosine phosphorylation, and importantly, abrogates IRS downstream PI3K-AKT signaling, thus promoting FOXO1-mediated gluconeogenesis, and finally causing insulin resistance (Figure 5A) [85,86,87]. Notably, both JNK1 and JNK2 can phosphorylate IRS1 [81], and JNK is also able to phosphorylate IRS2 [86,88,89].
Han et al. found that JNK activity in macrophages of adipose tissue contributes to obesity-induced insulin resistance and inflammation [90]. Mechanistically, the activated JNK1/2 promotes c-Jun, ATF-2, and ELK1 to bind to the promoters of pro-inflammatory cytokine-coding genes such as IL-6, subsequently causing the transcription of these genes, finally resulting in the secretion of pro-inflammatory cytokines from macrophages. Similarly, Perry et al. found that JNK signaling promotes the secretion of IL-6, causing an increase in FFAs that further facilitate the liver’s ability to synthesize glucose, finally resulting in insulin resistance [30]. Together, these studies indicate that JNK activation exacerbates the progression of T2D through the promotion of metabolic inflammation (Figure 5B). In regard to the regulation of the TSH–thyroid hormone axis by JNK, Belgardt et al. discovered that the expression levels of T3 and TSHβ in the pituitary were increased in HFD-feeding mice, whose JNK1 is deficient in the central nervous system, indicating that JNK inhibits the TSH–thyroid hormone axis. Moreover, body mass and liver triglyceride content were also reduced in these mice [91]. Similarly, Sabio et al. found that JNK1 deficiency in the CNS also caused a decrease in adiposity in HFD-feeding mice. Taken together, these studies showed that HFD-stimulated JNK1 promotes adiposity by negatively regulating the TSH–thyroid axis [92]. Moreover, the activated JNK promotes AP-1 transcription factors, which include c-Jun and c-Fos, to promote Dio2 gene transcription and expression. Then, Dio2 makes T4 convert into T3, which further inhibits TSHβ gene expression (Figure 5C) [93]. Subsequently, the reduced TSHβ results in a reduction of thyroid hormone and an increase in adiposity, which finally leads to insulin resistance. Notably, Vernia et al. found that HFD-feeding-activated JNK3 in hypothalamic neurons is needed for controlling appetite for food [94]. Interestingly, in contrast to control animals, HFD-feeding JNK3 knockout mice are apt to develop obesity and subsequent insulin resistance, suggesting neuron-specific JNK3 plays a role in improving insulin resistance [94].
Vernia et al. found that hepatocyte-specific JNK1/2 knock-out improved insulin sensitivity and significantly ameliorated fatty liver in a FGF21-dependent manner [95,96]. Mechanistically, JNK first stimulates c-JUN to bind to the promoter of the NcoR1 gene and promotes the transcription and subsequent expression of NCOR1. As a transcriptional corepressor, NCOR1 further represses the PPARα and RXR-mediated PPARα response genes, such as FGF21 expression, thus affecting fatty acid oxidation and ketogenesis, which finally result in insulin resistance (Figure 5D). In addition to these above mechanisms, JNK activity can affect insulin secretion from pancreatic β-cells. In the pancreatic β-cell, the elevated FFAs in plasma during obesity will result in the sustained activation of JNK, which, in turn, negatively regulates insulin secretion and finally causes β-cell dysfunction and death [68,97]. Collectively, these studies demonstrate that JNK plays an important role in obesity and type II diabetes.

3.1.3. JNK in Atherosclerosis

Atherosclerosis is a multi-step systemic inflammation and metabolic disease, which ultimately leads to myocardial infarction or stroke [36]. Both obesity and diabetes are known to contribute to atherosclerosis. Atherosclerosis onset and progression involve the following events: (1) An initiating stimulus, including vascular injury, hypercholesteremia, or chronic inflammation. These stimuli cause endothelial cell dysfunction and/or apoptosis, thus increasing vessel wall permeability to lipids and local inflammation. Under these conditions, endothelial cells then express cytokines that recruit monocytes and leukocytes to the area; (2) monocytes then trans-migrate across the blood vessel wall and differentiate into macrophages. As the vessel is lipid-laden, the macrophages ingest the low-density lipoproteins (LDLs), and then turn into foam cells. These foam cells further promote lesion progression by increasing local inflammation and ROS production; (3) under the actions of endothelial cell-secreted cytokines, T and B lymphocytes are recruited to the plaque. Smooth muscle cells are also recruited to the lumen and begin to proliferate and secrete collagen, elastin, and other extracellular matrix proteins [36]. Although multiple cell types contribute to this complex process, elevated JNK activity is observed in each of these steps [36]. As mentioned above, JNK1 or JNK2 plays an important role in inflammation and cytokine production. Moreover, JNK1 or JNK2 is expressed in all of the cell types relevant to the onset and progression of atherosclerosis: endothelial cells, smooth muscle cells, macrophages, and T cells, all of which are associated with the formation and development of atherosclerosis. Specifically, JNK1 promotes apoptosis in the endothelium after chronic inflammation and promotes atherosclerosis. Similar to the endothelium, JNK1 in bone marrow-derived immune cells (including monocytes) also promotes apoptosis after chronic inflammation, which leads to less atherosclerosis in mice. JNK2 knockout mice show less atherosclerosis through a reduced number of foam cell formation. These studies suggest that JNK is involved in inflammation and cytokine production, mediates the apoptosis of the endothelium, immune cells, and foam cells, and has different functions in different cells (macrophages, T cells, and endothelial cells). Previous studies have shown that JNK over-activation promotes the progression of atherosclerosis [36,98]. Furthermore, Ricci et al. found that JNK2 ablation can mitigate atherosclerosis progression, implying that JNK2 can promote atherosclerotic lesion formation [99]. However, Babaev et al. demonstrated that a hematopoietic cell-specific JNK1 deficiency promoted atherosclerosis in LDLR−/− mice. By contrast, Nofer et al. found that JNK promoted atherosclerosis onset. Similarly, Amini et al. found that JNK1 ablation protected atherosclerosis formation [100]. These studies suggest that JNK is critical for the onset and progression of atherosclerosis.

3.1.4. JNK in Non-alcoholic Fatty Liver Disease (NAFLD)

NAFLD includes non-alcoholic fatty liver (NAFL) and non-alcoholic steatohepatitis (NASH) [101]. Hypertension and cardiovascular diseases, hepatic gluconeogenesis-triggered hyperglycemia, and insulin resistance-caused hyperinsulinemia all lead to NAFL disease [101]. Although the mechanism is not fully understood, JNK activation in response to hepatic metabolic stresses, including inflammation, cytokines, de novo lipogenesis, and lipolysis, has been identified as a reason for NAFL/NASH [102,103,104,105,106].
It is well known that the liver contains about 60% parenchymal cells and over 30% non-parenchymal cells, including hepatic stellate cells. Among these cells, hepatocytes, a type of parenchymal cell, are the cornerstone of the liver. Hepatocytes express three JNK isoforms [107]. Diets, especially hyper-nutrition-type diets, including HFD, easily activate hepatic JNK [96]. Mechanically, HFD first activates JNK in the hepatocytes [108,109]. Then, the activated p-JNK further phosphorylates SAB, a mitochondrial outer membrane protein, resulting in mitochondrial ROS production, which further activates the ASK1-MKK4/7-JNK axis [107,110], thus forming a JNK-SAB-ROS-ASK1-JNK feedback loop, finally resulting in sustained JNK activation. This feedback loop is considered a critical mechanism in the development of hepatic steatosis [30,51]. The important role of JNK in NASH is proved by the following reports: ASK1 knockout mice are protected from hepatic steatosis [101,111]; MLK knockout mice also reduce the level of triglycerides [112]; similarly, JNK knockout mice avoid HFD-triggered NASH [113].

3.2. The Role of JNK Signaling in Neurological Diseases

JNK signaling exerts different roles in neurogenesis, axonal growth, axonal transport, and brain metabolism [114]. The deregulated JNK pathway leads to developmental axonal defects. Moreover, the over-activation of the JNK pathway, which means the kinase activity of JNK is constitutively increased, also contributes to CNS pathologies and motor neuron (MN) diseases [8]. Here, we briefly summarize the function of JNK in CNS pathologies such as glioblastoma progression, neurodegenerative diseases, and CNS regeneration/repair after an injury.

3.2.1. The JNK Pathway in Glioblastoma Progression

The JNK pathway is highly active in the central nervous system (CNS) as compared with other tissues. Previous studies revealed a dual role of JNK in cell death and survival, which is very important for glioblastoma (GB) tumorigenesis and neurodegeneration [115,116,117]. GB is a malignant brain tumor [114,118]. Over-activation of the JNK pathway is a hallmark related to glial cell proliferation and cancer stem cell-like properties [117,119,120]. Recent findings showed that JNK activity correlates with GB aggressiveness and that JNK promotes GB progression and infiltration [117,121]. Mechanically, JNK activation leads to upregulation of matrix metalloproteases (MMPs) [117], causing extracellular matrix degradation and promoting GB cell infiltration [117]. Interestingly, the progressive increase of JNK pathway activation in GB samples is associated with the localization of the Grnd receptor’s ligand in glial cells [121]. Recently, Jarabo et al. found that GB cells can secrete Impl2 to induce neuronal changes, contributing to tumor progression. Importantly, this process is also regulated by JNK activity [122]. Together, these studies indicate that GB cells use different ways to activate the JNK signaling pathway, further promoting GB progression.

3.2.2. The JNK Pathway in Neurodegenerative Diseases

Parkinson’s disease (PD), Alzheimer’s disease (AD), and Huntington’s disease (HD) are the three most common neurodegenerative diseases [26]. CNS-specific JNK3 has been shown to be a mediator for these neurodegenerative diseases. In AD patients, there are many diffuse plaques in cortical and hippocampal neurons [26,123]. These plaques mainly contain β-amyloid peptide (Aβ) and hyperphosphorylated Tau protein. Aβ induces neuronal apoptosis through JNK-dependent downregulation of Bcl-w, resulting in cortical neuron apoptosis [124]. JNK is also responsible for phosphorylating Tau to promote tangle formation [125]. In addition, JNK3 over-activation promotes AD progression as JNK3-deficient mice display synapse regrowth and a decrease in Aβ [114,126]. These mice are protected from Aβ-induced neuronal cell apoptosis through a JNK3-mediated AP-1-dependent FasL manner [123,124,127].
JNK also mediates PD progression. Choi et al. found that neurotoxin-induced JNK3 activation mediated dopaminergic neuron death in the substantia nigra [128]. JNK over-activation induces apoptosis of dopaminergic neurons [129]. Consistently, JNK3−/− and JNK2−/− mice prevented dopaminergic cell death induced by the neurotoxin MPTP [26]. Mechanistically, JNK-mediated autophagy and apoptosis play key roles in PD progression [129]. Bcl-2 has been identified as a critical protein with the ability to suppress autophagy and apoptosis through inhibiting Beclin-1 and Bax, respectively. Moreover, both JNK and P38 mediate dopaminergic neuron death in PD via apoptosis by increasing the Bax/Bcl-2 ratio. With regard to autophagy, JNK-mediated BCL-2 phosphorylation also suppresses the functions of Bcl-2 in autophagy. Additionally, recent findings indicate that receptor-interacting protein kinase 1 (RIPK1) is upregulated in PD in in vitro and in vivo models. RIPK1 promotes cell apoptosis and reactive oxygen species (ROS) production through the activation of the JNK pathway, thus leading to dopaminergic cell death. These discoveries suggest that JNK, together with other factors such as p38, BCL-2, Bax, Beclin-1, and RIPK1, mediate the progression of PD.
HD disease is caused by the degeneration of projection neurons in the striatum and the cerebral cortex. In HD patients, there are some amplifications of glutamine repeat (poly Q) in huntingtin protein (HTT), which trigger protein aggregation and neuronal degeneration [130]. During this process, polyQ-containing proteins activate JNKs. Furthermore, Morfini et al. demonstrated that aberrant HTT would impair axonal transport (FAT) through the JNK-mediated phosphorylation of the kinesin-1 protein [131]. Furthermore, JNK kinase inhibitors conferred neuroprotection in an HD mouse model [132], providing further evidence that JNK functions in HD progression. Taken together, these observations have demonstrated that the JNK signaling pathway is critical for the progress of neurodegenerative diseases.

3.2.3. JNK in Excitatory Toxicity of Hippocampal Neurons

The expression of JNK3 in the fore- and hind-brain regions is high. Normally, mice treated with kainic acid, a neurotoxic reagent, display significant apoptosis of hippocampal neurons [26]. JNK3 knockout mice also avoid glutamate-induced excitatory toxicity. Similarly, mice knocking in a mutant c-Jun (S63A and S73A) that cannot be phosphorylated by JNK display resistance against glutamate-induced neuronal toxicity [133]. These studies indicate that JNK3 activity promotes neuronal excitatory toxicity.

3.2.4. Abnormal Activation of the JNK Pathway in Multiple Sclerosis

Multiple sclerosis (MS) is a chronic demyelinating neurodegenerative disease of the CNS, manifesting as myelin sheath damage. Zhang et al. found that JNK was overactivated in chronic active MS plaques [134]. Bagnoud et al. also found that the phosphorylated JNK is increased during demyelination-triggered MS [135]. During the progression of MS, the phosphorylation of JNK by ASK1 plays critical roles in multiple processes, such as oligodendrocyte destruction, demyelination, neuroinflammation, immune dysregulation through T cell activation (T cell apoptosis), and oxidative damage. For instance, the TLR-ASK1-JNK pathway is active in glial cells and important for autoimmune demyelinating disorders [136]. In addition to the TLR (Toll-like receptor), other factors, including reactive oxygen species (ROS), oxidative stress, and inflammation, can activate ASK1 during MS progression [137]. These reports indicate that the over-activated ASK1-MKK4/7-JNK signaling axis plays a critical role in the pathogenesis of MS; that is, deregulated JNK signaling contributes to MS progression [134,135,137].

3.2.5. JNKs in CNS Damage Repair

Some factors, such as brain stroke and spinal cord injury, often trigger CNS damage. Because the self-repair or regenerative ability of this tissue is limited, the damaged CNS often displays permanent disabilities [114]. Previous studies have shown that JNK signaling has an important function in degeneration and the repair of nerve injuries. That is, retrograde transport of JNK in the injured axons changes the transcription of ATF3, thus affecting axon growth [138,139]. In addition, suppression of JNK signaling delays axonal degeneration [140]. DLK-JNK signaling also contributes to axonal regeneration after spinal cord injury [141,142,143,144]. Additionally, JNK3 interacts with KLF9 to inhibit axon regeneration upon nerve injury [145]. Overall, these reports clearly show the relevance of JNK to regeneration/repair after CNS injury.

3.3. JNK Pathway in Chronic Inflammatory Disease

Autoimmune disorders are a group of chronic inflammatory diseases, such as autoimmune arthritis (AA), osteoarthritis (OA), and inflammatory bowel disease (IBD). Here, we take AA, OA, and IBD as examples to demonstrate the function of JNK in chronic inflammatory disease.

3.3.1. The JNK Pathway in Autoimmune Arthritis

Rheumatoid arthritis (RA), psoriatic arthritis (PsA), and ankylosing spondylitis (AS) are three common autoimmune arthritis (AA) diseases [146]. The common characteristics of these diseases is the destruction of the joints and bones. Interestingly, extensive lymphocyte infiltration occurs in the inflamed joints in AA [146]. Also, pro-inflammatory cytokines, for instance, tumor necrosis factor-alpha (TNF-α), are significantly elevated in serum and joint fluid [146,147].
Fukushima et al. found that the phosphorylation of JNK was increased in the joint tissues of mice with inflammatory arthritis [148,149]. Similarly, JNK activation was also observed in synovial tissues from patients with psoriatic arthropathy and RA patients [150,151]. Recently, Li et al. found that JNK signaling was also involved in the pathogenesis of ankylosing spondylitis [152]. Mechanically, JNK is activated by pro-inflammatory cytokines, including TNF-α, causing autoimmune arthritis [146]. Additionally, JNK also regulates the expression of metalloproteases, promoting joint destruction [26], as JNK1/JNK2-deficient rheumatoid arthritis mice were protected from joint damage and reduced the expression level of metalloproteinase [26]. These studies clearly suggested that deregulated JNK signaling plays a role in autoimmune arthritis.

3.3.2. The JNK Pathway in Osteoarthritis

Osteoarthritis (OA) is a joint degeneration-related disease. The characteristics of OA mainly include several pathologic changes, such as joint inflammation, erosion of articular cartilage, and osteophyte formation [153]. JNK has been shown to aggravate this pathological process. For example, upon activation, JNK promotes cartilage destruction. Specifically, JNK is first activated by the pro-inflammatory cytokine TNFα; then, the activated JNK phosphorylates c-Jun, promoting MMP-13 (a cartilage-degrading enzyme) transcription and expression, which finally causes a decrease in proteoglycan synthesis. Additionally, excessive MMP-13 expression in chondrocytes contributes to cartilage degeneration [26,154]. Additionally, Yang et al. found that the elevated levels of CXCL8 and CXCL11 in the synovial fluids also aggravate OA progression by activating the JNK pathway, which mediates the apoptosis of chondrocytes [155].

3.3.3. JNK Functions in Chronic Inflammatory Bowel Disease

Crohn’s disease and ulcerative colitis are two chronic inflammatory bowel diseases (IBDs). Several pro-inflammatory cytokines in the intestinal microenvironments, for example, TNFα, can activate JNK signaling, which accelerates the development of IBD [156]. Previous studies showed that JNK was highly activated in the intestinal tissues from IBD patients [156,157]. Similarly, Mistsuyama et al. found that the active form of JNK was expressed in the nuclei of intestinal cells, macrophages, and lymphocytes, further providing indirect evidence that pro-inflammatory cytokines activate the JNK pathway in IBD [157]. Chromik et al. investigated the function of JNK isoforms in IBD and found that JNK2 deficiency aggravated DSS-induced colitis in mice; however, JNK1 deficiency did not completely block colitis development [158]. By contrast, the JNK inhibitor SP600125 partially protected against colonic injury [156].

3.4. Pro- and Anti-Oncogenic Roles of JNK in Cancer

The JNK pathway has a dual role in different types of cancers, including hepatocellular carcinoma (HCC), pancreatic cancer, prostate cancer, multiple myeloma and oral cancer, non-small-cell lung cancer (NSCLC), glioblastoma, papilloma, intestinal tumors, breast cancer, squamous cell carcinoma, Burkitt’s lymphoma, and ovarian cancer [1,9,26,57,114,159]. The dual effects of JNK in oncogenesis are summarized in Figure 6. On the one hand, JNK can promote cell death, eliminating pre-tumorigenic cells. On the other hand, it can also stimulate tumorigenesis [26,115].

3.4.1. The Dual Role of JNK in Hepatocellular Carcinoma (HCC), Multiple Myeloma, Prostate Cancer, and Oral Cancer

JNK has a tumor promoting or suppressing function in HCC, pancreatic cancer, multiple myeloma, and oral cancer (Figure 6A). JNK1 deficiency in hepatocytes and JNK2 deficiency in the liver promote HCC progress, indicating its role in inhibiting tumor development [159]. Similarly, JNK1 deficiency significantly decreases susceptibility to diethylnitrosamine-induced hepatocarcinogenesis [26]. However, JNK has an oncogenesis role in non-parenchymal cells of the liver through the production of pro-tumorigenic cytokines [159]. Hideshima et al. found that the JNK inhibitor SP600125 arrests MM cells in the cell cycle and causes cell growth inhibition through activation of the NF-κB pathway [160]. However, Sharkey et al. indicated that SP600125 abolished the apoptosis of multiple myeloma cells [161]. These two studies clearly showed the controversial role of JNK in multiple myeloma through different mechanisms affecting cell apoptosis and survival. In the same way, the JNK pathway plays dual functions in prostate cancer [9]. JNK has been shown to contribute to the apoptosis of prostate cancer through endoplasmic reticulum stress, the death receptor-dependent apoptotic pathway, and the mitochondria pathway [9]. By contrast, several studies showed that the JNK pathway is involved in prostate cancer progression [9,162,163]. For example, Sung et al. found that Jazf1 promoted prostate cancer progression by activating JNK signaling in DU145 prostate cancer cells [162]. Similarly, Du et al. discovered that CC chemokine receptor 7 (CCR7) activated the Notch1 pathway, resulting in a significant increase of phosphorylated JNK, which further promoted the migration of prostate cancer cells [163]. In regard to oral cancer, the JNK pathway has been shown to play an oncogenic or tumor-suppressive role through action alone or synergistically with other MAPKs [164]. For instance, Noutomi et al. found that after activation, JNK further enhanced TRAIL-induced apoptosis of oral cancer cells [165]. Kim et al. discovered that JNK, together with ERK, were implicated in ROS-induced apoptosis of OSCC cells [166]. These studies showed that JNK has a tumor-inhibiting function in oral cancer. On the contrary, Gross et al. found that inhibition of JNK suppresses tumor growth in oral cancer by downregulating IL-8 signaling, VEGF activity, and EGFR activation [167], suggesting JNK’s pro-oncogenic role in oral cancer.

3.4.2. The Pro-Oncogenic Role of JNK in NSCLC and Glioblastoma

Emerging evidence suggests that the sustained activation of JNK contributes to the progression of NSCLC and glioblastoma [159]. Since the pro-oncogenic role of JNKs in glioblastoma has been summarized in this review (in the section on JNK’s role in neurological diseases), here, we briefly discuss the role of JNK in NSCLC. Luo et al. found that LINC00958 promoted NSCLC cell proliferation by activating the JNK pathway [168]. Recently, Lin et al. reported that KIAA1429 promoted NSCLC tumorigenesis through the activation of the JNK pathway [169]. These studies showed a pro-oncogenic role of the JNK pathway in NSCLC (Figure 6B).

3.4.3. JNK Functions as a Tumor Suppressor in Intestinal Tumors, Papilloma, and Breast Cancer

The tumor-suppressive role of the JNK pathway in intestinal cancer, papilloma, and breast cancer is shown in Figure 6C. First, its role in breast cancer was supported by the findings of Cellurale and others that JNK1/2 ablation significantly increases tumor formation in breast cancer [170]. Subsequently, Shen et al. reported that cambogin induced the apoptosis of breast cancer cells by activating JNK signaling [171]. More recently, Itah et al. also found that JNK inhibition contributed to the progression of breast cancer [172]. Similarly, JNK also acts as a tumor suppressor in intestinal cancer and papilloma. For example, Tong et al. found that JNK1 knockout mice spontaneously developed intestinal cancer [173]. Recently, Kwak also reported that isolinderalactone exerted its anticancer effect on oxaliplatin-resistant colorectal cancer cells by inducing JNK-mediated apoptosis [174]. Choi et al. found that JNK1 can phosphorylate Myt1 to induce caspase-3-mediated apoptosis, thus preventing the formation of skin cancer, and that JNK1 knockout mice developed more UV-induced papilloma than JNK wild-type mice [175].

3.4.4. JNK Inactivation Suppresses Tumorigenesis in Ovarian Cancer, Skin Cancer, and Lymphoma

Emerging evidence showed that inhibition of JNK signaling prevented the oncogenesis of skin cancer, lymphoma, and ovarian cancer (Figure 6D) [159]. For instance, JNK2-KO mice exerted suppression of skin tumorigenesis [176], indicating the tumor-promoting role of JNK2 in skin cancer [57,176]. Additionally, Ding et al. found that targeting the JNK pathway suppressed Burkitt’s lymphoma cell proliferation [177]. Yang et al. found that JNK3 inactivation enhanced BH3 mimetic S1-induced apoptosis in cisplatin-resistant human ovarian cancer cells [178]. Similarly, JNK1 inhibition also suppressed ovarian cancer growth [179]. These results suggested the suppressive roles of different JNK isoforms in skin cancer, lymphoma, and ovarian cancer.

3.5. The JNK Pathway in Infectious Diseases

The JNK pathway has been shown to function in multiple infectious diseases, which are caused by pathogens, including viruses, bacteria, fungi, and parasites [61]. Thus, we briefly discuss JNK’s role in these infectious diseases.

3.5.1. JNK in Viral Diseases

The JNK pathway can be activated by viruses, which, in turn, regulates many viral infectious diseases [180]. The majority of viruses use their encoding proteins to activate JNK signaling. For instance, the NS1 protein of the influenza A virus can activate JNK [60]. Similarly, the Tat protein of HIV-1 activates JNK signaling through a Nox-mediated mechanism [62]. The activated JNK, generally speaking, further supports viral infection and replication [61]. Previous studies have shown that JNK inactivation results in significantly reduced replication of viruses, such as dengue virus [181], influenza virus [182], and veterinary viruses [180]. Mechanistically, JNK promotes apoptosis of infected host cells, accelerating viral infection [61]. In addition, JNK is also important for viral replication via autophagy. In addition to the contribution of JNK to the replication of most viruses, JNKs also negatively regulate the replication of the oncolytic vaccinia virus [61].

3.5.2. JNK in Bacterial Infections

Previous studies have shown that JNK can be activated upon bacterial infection in multiple mammalian cell types [40,183]. Correspondingly, bacteria have evolved several mechanisms for activating JNK. For example, lipopolysaccharide (LPS), an outer membrane protein of Gram-negative bacteria, can activate the JNK signaling pathway [184]. Nguyen et al. showed that pneumolysin, a virulence factor of Streptococcus pneumoniae, can activate the JNK pathways to induce ATF3 expression [185]. Similarly, Escherichia coli (E. coli)-produced shiga toxin 1 is able to activate JNK, and subsequently cause the apoptosis of intestinal epithelial cells [186]. In addition, some bacteria themselves can regulate host JNK activity through posttranslational modification of JNK proteins [61]. For instance, AvrA protein, a salmonella effector protein, can promote acetylation of specific host MAPKKs, resulting in suppression of JNK signaling and ultimately hampering the host immune response against salmonella [187].

3.5.3. The JNK Pathway in Fungal and Parasitic Infections

The JNK pathway is activated upon fungal infection [61]. For example, JNK in human bronchial alveolar epithelial cells is activated by the virulence factor gliotoxin when infected with the mold Aspergillus fumigatus. Then the activated JNK mediates the apoptosis of human bronchial alveolar epithelial cells, ultimately causing invasive aspergillosis [188]. Additionally, the JNK pathway is also activated in host antifungal responses. An example is that Candida albicans-activated JNK1 signaling has a negative effect on antifungal innate immunity [189].
Similar to bacterial or fungal infections, JNK can be activated by parasitic infections. For example, JNK2 is activated by Toxoplasma gondii during the infection process [190]. Then, the activated JNK2 plays a dual role in host resistance and immunity [191].

3.6. The JNK Pathway in Other Diseases

JNK also functions in other diseases. Here, we briefly discuss its role in other diseases that have not been mentioned above.

3.6.1. Hearing Loss

JNK is reported to be involved in deafness via promoting apoptosis [26]. Eshraghi et al. found that the JNK pathway is activated in hair cells under external stresses such as cochlear implantation trauma and noise. The activated JNK then induces hair cell apoptosis, ultimately causing deafness [192].

3.6.2. Ischemic/Reperfusion Injury

Ischemia/reperfusion injury often occurs in the brain, liver, heart, and kidneys. Although reperfusion is necessary for survival, it will result in tissue injury and activate inflammatory signaling, ultimately promoting JNK activation [36]. Taking myocardial infarction as an example, inflammatory signaling is stimulated in ischemic or necrotic myocardial cells, which ultimately leads to ventricular remodeling. This process involves ASK1-JNK1/2 signaling because ASK1 and JNK1/2 KO mice display reduced cardiac remodeling [36,193,194].

3.6.3. Cardiac Hypertrophy

Cardiac hypertrophy is defined as the thickening of the heart muscle under a variety of pathological stresses, such as mechanical stress and inflammation [36]. Although transient enlargement of the heart is a normal physiological phenomenon to adapt to these stimulations, the heart will suffer maladaptive hypertrophy under sustained pathological stresses, ultimately leading to heart failure [195]. The process of cardiac hypertrophy involves the activation of JNK signaling. For example, p-JNK is detected in patients with heart hypertrophy; therefore, JNK signaling is thought to promote myocyte growth, which marks pathologic hypertrophy [36].

3.6.4. Abdominal Aortic Aneurysms

Abdominal aortic aneurysms (AAAs) are one kind of arterial aneurysm. Chronic inflammation and vascular smooth muscle hypertrophy are considered the main reasons for AAA [36]. AAA formation involves JNK activation, as p-JNK is widely expressed in aneurysm tissue from patients with AAA [196]. Moreover, JNK expression in aneurysm tissue affects extracellular matrix metabolism, which further promotes the progression of AAA. Interestingly, pharmacological JNK inhibition alleviates AAA symptoms [196].

3.6.5. Renal Fibrosis

Renal fibrosis mainly refers to glomerular fibrosis and tubulointerstitial fibrosis, which ultimately develop into end-stage renal failure. These two processes involve activation of JNK signaling, as phosphorylated JNK is widely observed in glomerular cells and renal tissue from patients with acute and chronic renal injury [197]. Importantly, JNK inhibitors can ameliorate renal fibrosis. A key mechanism is that JNK can phosphorylate SMAD3, thus promoting the expression of pro-fibrotic genes, which further aggravate renal fibrosis [197].

3.6.6. Autosomal Dominant Polycystic Kidney Disease

Polycystic kidney disease (PKD) is a degenerative kidney disease wherein the renal tubules are filled with multiple fluid-filled cysts [198]. Autosomal dominant polycystic kidney disease (ADPKD) is one kind of PKD. ADPKD has been shown to be caused by mutations of the polycystin-1 (Pkd1) and Pkd2 genes. Smith et al. recently found that the JNK pathway is involved in ADPKD as inhibition of JNK activity reduces cyst growth [198].

4. Targeting the JNK Pathway as a Therapeutic Strategy for Treating Human Diseases

Since the deregulated JNK pathway is widely implicated in multiple types of diseases, it is reasonable that pharmacological inhibition of JNK activation or JNK activity is an effective strategy to combat or ameliorate the disease conditions. Currently, JNK inhibitors are used as candidate drugs for ameliorating deregulated JNK signaling-related diseases. Here, we briefly summarize the potential applications of JNK inhibitors in treating human diseases.

4.1. Overview of JNK Inhibitors

Broadly, JNK inhibitors mainly include the following several types of inhibitors: ATP-competitive inhibitors, ATP non-competitive inhibitors, and small peptide inhibitors. Mechanistically, ATP-competitive inhibitors interact with the hinge region of the ATP binding site. The non-competitive inhibitors do not bind to ATP-binding sites but bind to the catalytic sites. By contrast, small peptide inhibitors regulate JNK kinase activity by interacting with JNK’s scaffold or its upstream/downstream molecules or by changing the subcellular localization of JNK [27]. Additionally, some natural phytochemicals also exert inhibiting effects against JNK activity.

4.2. The Application of Synthetic JNK Inhibitors in Human Diseases

Up to now, there are a variety of JNK inhibitors that have been studied for ameliorating the disease state. These inhibitors and their applications for treating diseases are summarized in Table 1. In this review, we take several classic JNK inhibitors as examples to elucidate the application of these inhibitors in treating human diseases.

4.2.1. ATP-Competitive Inhibitors in Human Diseases and Their Action Mechanisms

Here, we give a brief review of the application of three classic ATP-competitive inhibitors in human diseases. The ATP-competitive inhibitors that were used for therapeutic intervention mainly include SP600125, CEP-1347, and AS601245. The three JNK inhibitors occupy an ATP-binding site. SP600125 directly inhibits JNK activity, whereas CEP-1347 inhibits MLK to weaken JNK signaling [68,204,205]. SP600125, a compound of anthrapyrazolone, is the most extensively studied ATP-competitive JNK inhibitor [1]. Although it lacks specificity, SP600125 has been shown to have therapeutic potential. SP600125 was used to study recovery from a number of the deregulated JNK pathway-associated diseases such as ischemia/reperfusion, cancer cell apoptosis, T2D, acute kidney injury, inflammation, viral infections, and sepsis-induced lung injury [27,197,206]. Mechanistically, as a reversible ATP-competitive inhibitor, SP600125 competitively interacts with the ATP binding site of JNK, leading to a reduction of JNK phosphorylation (reduction of JNK activity), either repressing nuclear transcription factors-mediated transcription of target genes or inhibiting phosphorylation of non-nuclear proteins [207]. Because of the important roles of these proteins in cell growth, proliferation, apoptosis, and metabolic stress, SP600125 can ameliorate those above-mentioned diseases. Additionally, SP600125 has been shown to inhibit the expression of inflammatory genes such as COX-2, IL-2, IFN-γ, and TNF-α, so it can also be used to improve inflammation [207]. As mentioned above, different diseases involve different JNK isoforms. For example, JNK3 is mainly involved in CNS-related diseases because of its brain-specificity. Moreover, not all isoforms of JNK play negative roles in disease progression. Thus, the non-specific inhibiting effects of SP600125 on JNK isoforms limit its application. CEP-1347, a semisynthetic derivative of the naturally occurring indolocarbazole (K-252a), has broad serine/threonine and tyrosine (trkA) kinase inhibitory activity and is used in pre-clinical studies for ameliorating Parkinson’s disease [68,208]. In terms of the action mechanism, CEP-1347 occupies an ATP-binding site of MLK, the MAP3K upstream of JNK. Binding of the ATP-binding site prevents phosphorylation and activation of MLK, thus negatively regulating JNK signaling [68]. AS601245, also known as bentamapimod, directly inhibits JNK3 at the ATP-binding site, resulting in inhibition of JNK3 activity [68]. Due to the role of JNK3 in neuronal death, AS601245 displays a significant neuroprotective effect against ischemia-induced neuronal cell death and cancer [1]. Additionally, it also exerted anti-inflammatory activity against rheumatoid arthritis [209].

4.2.2. ATP-Non-competitive Inhibitors in Human Diseases and Their Action Mechanisms

Next, we take several ATP-non-competitive JNK inhibitors (BI-78D3, BI87G3, and compound 9) as examples to elucidate the application of this type of inhibitor. BI-78D3 is the first ATP-non-competitive JNK inhibitor that was developed by Pellecchia and colleagues. Mechanistically, BI-78D3 inhibits JNK activity by interfering with the interaction of JNK1 and the JIP-derived peptide [29,199]. Functionally, BI-78D3 exerts protective effects against liver damage induced by concanavalin A [199]. Also, BI-78D3 was shown to restore insulin sensitivity [199]. Based on the structure–activity relationship, researchers have developed several JNK inhibitors. These inhibitors have similar action mechanisms; that is, they inhibit JNK signaling by blocking the JNK–JIP interaction [82,83]. For example, BI-87G3, a benzothiazole derivative, inhibits JNK activity by competing with JIP for binding to JNK [210]. Similarly, the thiadizole compound 9 competes with JIP-derived peptides to bind to JNK, thus inhibiting JNK activity. Compound 9 can enhance insulin sensitivity in a T2D murine model [211]. In addition, two types of JNK inhibitors were developed: JIP site inhibitors that target the JIP–JNK interaction and dual site inhibitors that not only target the JIP/JNK site but also disrupt the interaction of ATP and JNK. However, the therapeutic potential of these compounds is unknown [29].

4.2.3. Small Peptide Inhibitors in Human Diseases and Their Action Mechanisms

As discussed above, both ATP-competitive and ATP-non-competitive inhibitors have possible off-target effects. Small peptide inhibitors may provide a more targeted inhibition as they specifically target the JNK binding domain (JBD) or regulatory region [27]. One well-known JNK peptide inhibitor is the D-JNKI-1 peptide (also known as XG-102 or AM-111), which is synthesized according to the JIP and HIV TAT sequences. As for the action mechanism, the D-JNKI-1 peptide works by preventing the interaction of JNK with its JBD-dependent targets. For instance, D-JNKI-1 peptide prevents the activity of MKK4 and MKK7, the two JNK upstream MAPKKs, by interacting with the JBD homology domain [212]. D-JNKI-1 can penetrate the cells. D-JNKI-1 has protective effects against nerve injury [204] and can improve lipid metabolism and ameliorate T2D conditions [213]. Recently, D-JNKI-1 has been shown to have beneficial effects on inflammation, pain post-surgery, and hearing loss [27,68,200]. Islet Brain (IB1/2) peptide, another well-known JNK peptide inhibitor, contains a similar JBD domain. Mechanistically, the IB1/2 peptide resembles JNK-interacting protein (JIP) to inhibit JNK. IB1/2 can improve insulin resistance and glucose tolerance [214]. Other peptide inhibitors, whose structures are similar to those of JIP, all exert neuroprotective effects [215].

4.2.4. Dual ATP and Substrate-Competitive Kinase Inhibitors

Notably, designing dual ATP and substrate-competitive kinase inhibitors is a useful approach to exploring novel JNK inhibitors. Stebbins et al. used this strategy to design JNK inhibitors and obtained a bidentate compound (compound 19) that showed a beneficial effect against diabetes. Compound 19 inhibited JNK activity by concurrently targeting the ATP-binding site and substrate-binding sites [216].

4.3. The Application of Natural Phytochemicals That Inhibit JNK Activity in Human Diseases

Many natural phytochemicals showed inhibitory effects on JNK activity and were considered candidate drugs for therapeutic intervention against JNK-related diseases. For instance, the C66 natural variant of curcumin can alleviate diabetes and its complications by inhibiting JNK [27,201,202,203]. Mechanistically, C66 ameliorates diabetes-induced pathological changes by inhibiting JNK2 activity and upregulating Nrf2 expression [201]. Leupeol, a natural pentacyclic triterpene, can inhibit JNK activity. Previous studies have shown that leupeol has anti-hyperglycemic and anti-dyslipidemic activity and has a protective effect against lipopolysaccharide (LPS)-induced neuro-inflammation [217,218]. Leupeol inhibits LPS-induced neuroinflammation by inhibiting phosphorylation of JNK and decreasing the generation of pro-inflammatory cytokines, including TNF-α, iNOS, and IL-1β [217]. Gingerol can modulate lipid metabolism by inhibiting TNF-α-induced JNK activation [219]. These studies demonstrate that plant-based natural products that inhibit JNK activity also have therapeutic potential for human diseases.

4.4. The Challenges We Faced When We Were Developing JNK Inhibitors

Although small-molecule inhibitors of JNKs show beneficial efficacy against various JNK-related diseases to some extent, it should be noted that the non-specificity of these inhibitors should be considered. That is, when developing JNK inhibitors, off-target effects and potential side effects need to be considered. Here, in order to emphasize the limitations of current JNK inhibitors and help develop new JNK inhibitors in the future, we provide brief information about the (isoform) specificity and off-target effects of several above-mentioned JNK inhibitors, as well as the structural information of JNKs complexed with their corresponding inhibitors (Table 2).

5. Conclusions and Perspectives

The JNK signaling pathway regulates a variety of cell behaviors, such as cell proliferation, differentiation, and cell survival. Dysregulated JNK signaling is implicated in various human diseases such as metabolic disorders (e.g., obesity, T2D, and NAFLD), neurological disease (e.g., neurodegenerative disease), chronic inflammatory disease (e.g., RA and IBD), infectious disease, and cancer. Although we have made some progress in understanding the critical role of the JNK pathway in human disease, our knowledge about the function of JNK is still limited. Especially several questions, for instance, how the JNK pathway is strictly regulated under physiological and pathophysiological conditions and why the isoform-specific and cell type-specific responses of JNK are important for disease progress, remain to be addressed. Future studies are needed to elucidate the JNK isoform-specific role in a tissue-type-specific manner and to better explore the function of other molecules in the JNK pathway in disease progression.
Since JNKs play important roles in almost all human diseases, a common strategy for treating the disease with the target JNK, theoretically, would be desirable. For example, as described in Table 1, the JNK inhibitor SP600125 can be used to treat several diseases, such as ischemia/reperfusion, cancer, inflammation, viral infections, and sepsis-induced lung injury. However, this would require the identification of common or individual disorders of JNK signaling that occur under the corresponding disease states. Thus, identification of common or individual deregulated JNK signaling may be a future direction.
In terms of developing JNK inhibitors, off-target effects and potential side effects need to be considered. Small-molecule inhibitors such as ATP-competitive inhibitors, ATP non-competitive inhibitors, and small peptide inhibitors are extensively investigated, and some of these inhibitors show beneficial efficacy against various JNK-related diseases. However, avoiding the non-specificity of these inhibitors is still a great challenge. For example, because of the high homology (98% similarity) in the ATP binding pocket among JNK1, JNK2, and JNK3, it is a challenge to explore isoform-selective inhibitors. Therefore, it is urgent and necessary for us to identify substrate-specific or JNK isoform-specific inhibitors. In the future, targeting JNK kinase regulatory sites may be a new approach for discovering JNK isoform-specific inhibitors. In addition, because JNK signaling sometimes plays a conflicting role in some diseases, we need to consider and balance the advantages and side effects of JNK inhibitors. Thus, caution is still needed for the use of these JNK inhibitors in future clinical applications.

Author Contributions

Z.Y. and D.L. designed the manuscript. H.Y., L.H. and J.Y. wrote and revised the manuscript. H.Y. prepared figures of the manuscript. L.H. made the tables in the manuscript. J.Y. contributed to making the figures and tables in the manuscript. Z.Y. and D.L. made significant revisions and proofread the manuscript. Z.Y. and D.L. also provided support for the publication of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work received no external funding.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

MAPK: mitogen-activated protein kinase; JNK, c-Jun N-terminal kinase; ERK, extracellular signal-regulated kinase; SAPK, stress-activated MAP kinase; MLK, mixed-lineage protein kinase; ASK, apoptosis signal-regulating kinase; JIP, JNK-interacting protein; IKAP, the IκB kinase complex-associated protein; ACID, amyloid precursor protein intracellular domain; Notch-IC, Notch intracellular domain; DUSP1, dual specificity protein phosphatase; Sab, SH3 homology-associated BTK-binding protein; HPK, hematopoietic progenitor kinase; GCK, STE20 protein homologue germinal center kinase; NIK, Nck interacting kinase; ATF2, activating transcription factor 2; Elk1, ETS like-1 protein; HFD, high-fat diet; FFA, free fatty acid; T2D, type II diabetes; IRS, insulin receptor substrate; PI3K, phosphatidylinositide 3 kinase; TSH, thyroid-stimulating hormone; T3, triiodothyronine; T4, thyroxine; Dio2, type-2 deiodinase; CMS, cardiometabolic syndrome; LDL, low-density lipoprotein; LDLR, low-density lipoprotein receptor; NASH, Non-alcoholic steatohepatitis; CNS, central nervous system; SAB, SH3 homology-associated BTK-binding protein; MN, motor neuron; HTT, huntingtin protein; PD, Parkinson’s disease; AD, Alzheimer’s disease; HD, Huntington’s disease; KLF9, Kluppel-like transcription factor 9; OSCC, oral squamous cell carcinoma; EAE, experimental autoimmune encephalomyelitis; AAA, abdominal aortic aneurysm; and ADPKD, autosomal dominant polycystic kidney disease.

References

  1. Davies, C.; Tournier, C. Exploring the function of the JNK (c-Jun N-terminal kinase) signalling pathway in physiological and pathological processes to design novel therapeutic strategies. Biochem. Soc. Trans. 2012, 40, 85–89. [Google Scholar] [CrossRef] [PubMed]
  2. Guo, Y.J.; Pan, W.W.; Liu, S.B.; Shen, Z.F.; Xu, Y.; Hu, L.L. ERK/MAPK signalling pathway and tumorigenesis. Exp. Ther. Med. 2020, 19, 1997–2007. [Google Scholar] [CrossRef] [PubMed]
  3. Keshet, Y.; Seger, R. The MAP kinase signaling cascades: A system of hundreds of components regulates a diverse array of physiological functions. Methods Mol. Biol. 2010, 661, 3–38. [Google Scholar] [CrossRef]
  4. Plotnikov, A.; Zehorai, E.; Procaccia, S.; Seger, R. The MAPK cascades: Signaling components, nuclear roles and mechanisms of nuclear translocation. Biochim. Biophys. Acta 2011, 1813, 1619–1633. [Google Scholar] [CrossRef] [PubMed]
  5. Kim, E.K.; Choi, E.J. Pathological roles of MAPK signaling pathways in human diseases. Biochim. Biophys. Acta 2010, 1802, 396–405. [Google Scholar] [CrossRef]
  6. Pulverer, B.J.; Kyriakis, J.M.; Avruch, J.; Nikolakaki, E.; Woodgett, J.R. Phosphorylation of c-jun mediated by MAP kinases. Nature 1991, 353, 670–674. [Google Scholar] [CrossRef]
  7. Smeal, T.; Binetruy, B.; Mercola, D.A.; Birrer, M.; Karin, M. Oncogenic and transcriptional cooperation with Ha-Ras requires phosphorylation of c-Jun on serines 63 and 73. Nature 1991, 354, 494–496. [Google Scholar] [CrossRef]
  8. Schellino, R.; Boido, M.; Vercelli, A. JNK Signaling Pathway Involvement in Spinal Cord Neuron Development and Death. Cells 2019, 8, 1576. [Google Scholar] [CrossRef] [PubMed]
  9. Xu, R.; Hu, J. The role of JNK in prostate cancer progression and therapeutic strategies. Biomed. Pharmacother. 2020, 121, 109679. [Google Scholar] [CrossRef] [PubMed]
  10. Seki, E.; Brenner, D.A.; Karin, M. A liver full of JNK: Signaling in regulation of cell function and disease pathogenesis, and clinical approaches. Gastroenterology 2012, 143, 307–320. [Google Scholar] [CrossRef]
  11. Papa, S.; Choy, P.M.; Bubici, C. The ERK and JNK pathways in the regulation of metabolic reprogramming. Oncogene 2019, 38, 2223–2240. [Google Scholar] [CrossRef] [PubMed]
  12. Behrens, A.; Sibilia, M.; Wagner, E.F. Amino-terminal phosphorylation of c-Jun regulates stress-induced apoptosis and cellular proliferation. Nat. Genet. 1999, 21, 326–329. [Google Scholar] [CrossRef] [PubMed]
  13. Ventura, J.J.; Hubner, A.; Zhang, C.; Flavell, R.A.; Shokat, K.M.; Davis, R.J. Chemical genetic analysis of the time course of signal transduction by JNK. Mol. Cell 2006, 21, 701–710. [Google Scholar] [CrossRef] [PubMed]
  14. Kyriakis, J.M.; Brautigan, D.L.; Ingebritsen, T.S.; Avruch, J. pp54 microtubule-associated protein-2 kinase requires both tyrosine and serine/threonine phosphorylation for activity. J. Biol. Chem. 1991, 266, 10043–10046. [Google Scholar] [CrossRef] [PubMed]
  15. Kyriakis, J.M.; Avruch, J. pp54 microtubule-associated protein 2 kinase. A novel serine/threonine protein kinase regulated by phosphorylation and stimulated by poly-L-lysine. J. Biol. Chem. 1990, 265, 17355–17363. [Google Scholar] [CrossRef]
  16. Hibi, M.; Lin, A.; Smeal, T.; Minden, A.; Karin, M. Identification of an oncoprotein- and UV-responsive protein kinase that binds and potentiates the c-Jun activation domain. Genes. Dev. 1993, 7, 2135–2148. [Google Scholar] [CrossRef]
  17. Derijard, B.; Hibi, M.; Wu, I.H.; Barrett, T.; Su, B.; Deng, T.; Karin, M.; Davis, R.J. JNK1: A protein kinase stimulated by UV light and Ha-Ras that binds and phosphorylates the c-Jun activation domain. Cell 1994, 76, 1025–1037. [Google Scholar] [CrossRef]
  18. Kyriakis, J.M.; Banerjee, P.; Nikolakaki, E.; Dai, T.; Rubie, E.A.; Ahmad, M.F.; Avruch, J.; Woodgett, J.R. The stress-activated protein kinase subfamily of c-Jun kinases. Nature 1994, 369, 156–160. [Google Scholar] [CrossRef]
  19. Kallunki, T.; Su, B.; Tsigelny, I.; Sluss, H.K.; Derijard, B.; Moore, G.; Davis, R.; Karin, M. JNK2 contains a specificity-determining region responsible for efficient c-Jun binding and phosphorylation. Genes. Dev. 1994, 8, 2996–3007. [Google Scholar] [CrossRef] [PubMed]
  20. Heo, Y.S.; Kim, S.K.; Seo, C.I.; Kim, Y.K.; Sung, B.J.; Lee, H.S.; Lee, J.I.; Park, S.Y.; Kim, J.H.; Hwang, K.Y.; et al. Structural basis for the selective inhibition of JNK1 by the scaffolding protein JIP1 and SP600125. EMBO J. 2004, 23, 2185–2195. [Google Scholar] [CrossRef] [PubMed]
  21. Shaw, D.; Wang, S.M.; Villasenor, A.G.; Tsing, S.; Walter, D.; Browner, M.F.; Barnett, J.; Kuglstatter, A. The crystal structure of JNK2 reveals conformational flexibility in the MAP kinase insert and indicates its involvement in the regulation of catalytic activity. J. Mol. Biol. 2008, 383, 885–893. [Google Scholar] [CrossRef]
  22. Xie, X.; Gu, Y.; Fox, T.; Coll, J.T.; Fleming, M.A.; Markland, W.; Caron, P.R.; Wilson, K.P.; Su, M.S. Crystal structure of JNK3: A kinase implicated in neuronal apoptosis. Structure 1998, 6, 983–991. [Google Scholar] [CrossRef]
  23. Whitmarsh, A.J.; Cavanagh, J.; Tournier, C.; Yasuda, J.; Davis, R.J. A mammalian scaffold complex that selectively mediates MAP kinase activation. Science 1998, 281, 1671–1674. [Google Scholar] [CrossRef]
  24. Yasuda, J.; Whitmarsh, A.J.; Cavanagh, J.; Sharma, M.; Davis, R.J. The JIP group of mitogen-activated protein kinase scaffold proteins. Mol. Cell Biol. 1999, 19, 7245–7254. [Google Scholar] [CrossRef]
  25. Ait-Hamlat, A.; Zea, D.; Labeeuw, A.; Polit, L.; Richard, H.; Laine, E. Transcripts’ Evolutionary History and Structural Dynamics Give Mechanistic Insights into the Functional Diversity of the JNK Family. J. Mol. Biol. 2020, 432, 2121–2140. [Google Scholar] [CrossRef]
  26. Johnson, G.L.; Nakamura, K. The c-jun kinase/stress-activated pathway: Regulation, function and role in human disease. Biochim. Biophys. Acta 2007, 1773, 1341–1348. [Google Scholar] [CrossRef] [PubMed]
  27. Garg, R.; Kumariya, S.; Katekar, R.; Verma, S.; Goand, U.K.; Gayen, J.R. JNK signaling pathway in metabolic disorders: An emerging therapeutic target. Eur. J. Pharmacol. 2021, 901, 174079. [Google Scholar] [CrossRef] [PubMed]
  28. Castro-Torres, R.D.; Busquets, O.; Parcerisas, A.; Verdaguer, E.; Olloquequi, J.; Ettcheto, M.; Beas-Zarate, C.; Folch, J.; Camins, A.; Auladell, C. Involvement of JNK1 in Neuronal Polarization During Brain Development. Cells 2020, 9, 1897. [Google Scholar] [CrossRef]
  29. Bogoyevitch, M.A.; Ngoei, K.R.; Zhao, T.T.; Yeap, Y.Y.; Ng, D.C. c-Jun N-terminal kinase (JNK) signaling: Recent advances and challenges. Biochim. Biophys. Acta 2010, 1804, 463–475. [Google Scholar] [CrossRef] [PubMed]
  30. Maik-Rachline, G.; Wortzel, I.; Seger, R. Alternative Splicing of MAPKs in the Regulation of Signaling Specificity. Cells 2021, 10, 3466. [Google Scholar] [CrossRef] [PubMed]
  31. Barr, R.; Bogoyevitch, M. The c-Jun N-terminal protein kinase family of mitogen-activated protein kinases (JNK MAPKs). Int. J. Biochem. Cell Biol. 2001, 33, 1047–1063. [Google Scholar] [CrossRef] [PubMed]
  32. Gupta, S.; Barrett, T.; Whitmarsh, A.J.; Cavanagh, J.; Sluss, H.K.; Derijard, B.; Davis, R.J. Selective interaction of JNK protein kinase isoforms with transcription factors. EMBO J. 1996, 15, 2760–2770. [Google Scholar] [CrossRef]
  33. Nakano, R.; Nakayama, T.; Sugiya, H. Biological Properties of JNK3 and Its Function in Neurons, Astrocytes, Pancreatic beta-Cells and Cardiovascular Cells. Cells 2020, 9, 1802. [Google Scholar] [CrossRef] [PubMed]
  34. Vallerie, S.N.; Hotamisligil, G.S. The role of JNK proteins in metabolism. Sci. Transl. Med. 2010, 2, 60rv65. [Google Scholar] [CrossRef]
  35. Castro-Torres, R.D.; Landa, J.; Rabaza, M.; Busquets, O.; Olloquequi, J.; Ettcheto, M.; Beas-Zarate, C.; Folch, J.; Camins, A.; Auladell, C.; et al. JNK Isoforms Are Involved in the Control of Adult Hippocampal Neurogenesis in Mice, Both in Physiological Conditions and in an Experimental Model of Temporal Lobe Epilepsy. Mol. Neurobiol. 2019, 56, 5856–5865. [Google Scholar] [CrossRef] [PubMed]
  36. Craige, S.M.; Chen, K.; Blanton, R.M.; Keaney, J.F., Jr.; Kant, S. JNK and cardiometabolic dysfunction. Biosci. Rep. 2019, 39, BSR20190267. [Google Scholar] [CrossRef]
  37. Wang, J.; Tang, R.; Lv, M.; Wang, Q.; Zhang, X.; Guo, Y.; Chang, H.; Qiao, C.; Xiao, H.; Li, X.; et al. Defective anchoring of JNK1 in the cytoplasm by MKK7 in Jurkat cells is associated with resistance to Fas-mediated apoptosis. Mol. Biol. Cell 2011, 22, 117–127. [Google Scholar] [CrossRef]
  38. Hu, D.; Bi, X.; Fang, W.; Han, A.; Yang, W. GSK3beta is involved in JNK2-mediated beta-catenin inhibition. PLoS ONE 2009, 4, e6640. [Google Scholar] [CrossRef]
  39. Song, X.; Raman, D.; Gurevich, E.; Vishnivetskiy, S.; Gurevich, V. Visual and both non-visual arrestins in their “inactive” conformation bind JNK3 and Mdm2 and relocalize them from the nucleus to the cytoplasm. J. Biol. Chem. 2006, 281, 21491–21499. [Google Scholar] [CrossRef]
  40. Davis, R.J. Signal transduction by the JNK group of MAP kinases. Cell 2000, 103, 239–252. [Google Scholar] [CrossRef]
  41. Haeusgen, W.; Herdegen, T.; Waetzig, V. The bottleneck of JNK signaling: Molecular and functional characteristics of MKK4 and MKK7. Eur. J. Cell Biol. 2011, 90, 536–544. [Google Scholar] [CrossRef]
  42. Kragelj, J.; Palencia, A.; Nanao, M.H.; Maurin, D.; Bouvignies, G.; Blackledge, M.; Jensen, M.R. Structure and dynamics of the MKK7-JNK signaling complex. Proc. Natl. Acad. Sci. USA 2015, 112, 3409–3414. [Google Scholar] [CrossRef]
  43. Kirsch, K.; Zeke, A.; Toke, O.; Sok, P.; Sethi, A.; Sebo, A.; Kumar, G.S.; Egri, P.; Poti, A.L.; Gooley, P.; et al. Co-regulation of the transcription controlling ATF2 phosphoswitch by JNK and p38. Nat. Commun. 2020, 11, 5769. [Google Scholar] [CrossRef]
  44. Lu, W.; Liu, Y.; Gao, Y.; Geng, Q.; Gurbani, D.; Li, L.; Ficarro, S.B.; Meyer, C.J.; Sinha, D.; You, I.; et al. Development of a Covalent Inhibitor of c-Jun N-Terminal Protein Kinase (JNK) 2/3 with Selectivity over JNK1. J. Med. Chem. 2023, 66, 3356–3371. [Google Scholar] [CrossRef]
  45. Feng, Y.; Park, H.; Bauer, L.; Ryu, J.C.; Yoon, S.O. Thiophene-Pyrazolourea Derivatives as Potent, Orally Bioavailable, and Isoform-Selective JNK3 Inhibitors. ACS Med. Chem. Lett. 2021, 12, 24–29. [Google Scholar] [CrossRef]
  46. Coffey, E.T. Nuclear and cytosolic JNK signalling in neurons. Nat. Rev. Neurosci. 2014, 15, 285–299. [Google Scholar] [CrossRef] [PubMed]
  47. Nolen, B.; Taylor, S.; Ghosh, G. Regulation of protein kinases; controlling activity through activation segment conformation. Mol. Cell 2004, 15, 661–675. [Google Scholar] [CrossRef] [PubMed]
  48. Luhovy, A.; Jaberi, A.; Papillon, J.; Guillemette, J.; Cybulsky, A. Regulation of the Ste20-like kinase, SLK: Involvement of activation segment phosphorylation. J. Biol. Chem. 2012, 287, 5446–5458. [Google Scholar] [CrossRef] [PubMed]
  49. Kishimoto, H.; Nakagawa, K.; Watanabe, T.; Kitagawa, D.; Momose, H.; Seo, J.; Nishitai, G.; Shimizu, N.; Ohata, S.; Tanemura, S.; et al. Different properties of SEK1 and MKK7 in dual phosphorylation of stress-induced activated protein kinase SAPK/JNK in embryonic stem cells. J. Biol. Chem. 2003, 278, 16595–16601. [Google Scholar] [CrossRef] [PubMed]
  50. Fleming, Y.; Armstrong, C.; Morrice, N.; Paterson, A.; Goedert, M.; Cohen, P. Synergistic activation of stress-activated protein kinase 1/c-Jun N-terminal kinase (SAPK1/JNK) isoforms by mitogen-activated protein kinase kinase 4 (MKK4) and MKK7. Biochem. J. 2000, 352, 145–154. [Google Scholar] [CrossRef]
  51. Schattenberg, J.; Czaja, M. Regulation of the effects of CYP2E1-induced oxidative stress by JNK signaling. Redox Biol. 2014, 3, 7–15. [Google Scholar] [CrossRef]
  52. Salort, G.; Hernández-Hernández, E.; García-Fuster, M.; García-Sevilla, J. Regulation of cannabinoid CB and CB receptors, neuroprotective mTOR and pro-apoptotic JNK1/2 kinases in postmortem prefrontal cortex of subjects with major depressive disorder. J. Affect. Disord. 2020, 276, 626–635. [Google Scholar] [CrossRef] [PubMed]
  53. Bildik, G.; Akin, N.; Senbabaoglu, F.; Esmalian, Y.; Sahin, G.; Urman, D.; Karahuseyinoglu, S.; Ince, U.; Palaoglu, E.; Taskiran, C.; et al. Endogenous c-Jun N-terminal kinase (JNK) activity marks the boundary between normal and malignant granulosa cells. Cell Death Dis. 2018, 9, 421. [Google Scholar] [CrossRef]
  54. Waudby, C.; Alvarez-Teijeiro, S.; Josue Ruiz, E.; Suppinger, S.; Pinotsis, N.; Brown, P.; Behrens, A.; Christodoulou, J.; Mylona, A. An intrinsic temporal order of c-JUN N-terminal phosphorylation regulates its activity by orchestrating co-factor recruitment. Nat. Commun. 2022, 13, 6133. [Google Scholar] [CrossRef]
  55. Enomoto, A.; Suzuki, N.; Morita, A.; Ito, M.; Liu, C.; Matsumoto, Y.; Yoshioka, K.; Shiba, T.; Hosoi, Y. Caspase-mediated cleavage of JNK during stress-induced apoptosis. Biochem. Biophys. Res. Commun. 2003, 306, 837–842. [Google Scholar] [CrossRef]
  56. Yue, J.; Ben Messaoud, N.; López, J. Hyperosmotic Shock Engages Two Positive Feedback Loops through Caspase-3-dependent Proteolysis of JNK1-2 and Bid. J. Biol. Chem. 2015, 290, 30375–30389. [Google Scholar] [CrossRef]
  57. Hammouda, M.B.; Ford, A.E.; Liu, Y.; Zhang, J.Y. The JNK Signaling Pathway in Inflammatory Skin Disorders and Cancer. Cells 2020, 9, 857. [Google Scholar] [CrossRef] [PubMed]
  58. Morrison, D.K.; Davis, R.J. Regulation of MAP kinase signaling modules by scaffold proteins in mammals. Annu. Rev. Cell Dev. Biol. 2003, 19, 91–118. [Google Scholar] [CrossRef] [PubMed]
  59. Anfinogenova, N.D.; Quinn, M.T.; Schepetkin, I.A.; Atochin, D.N. Alarmins and c-Jun N-Terminal Kinase (JNK) Signaling in Neuroinflammation. Cells 2020, 9, 2350. [Google Scholar] [CrossRef]
  60. Nacken, W.; Wixler, V.; Ehrhardt, C.; Ludwig, S. Influenza A virus NS1 protein-induced JNK activation and apoptosis are not functionally linked. Cell Microbiol. 2017, 19, e12721. [Google Scholar] [CrossRef]
  61. Chen, J.; Ye, C.; Wan, C.; Li, G.; Peng, L.; Peng, Y.; Fang, R. The Roles of c-Jun N-Terminal Kinase (JNK) in Infectious Diseases. Int. J. Mol. Sci. 2021, 22, 9640. [Google Scholar] [CrossRef]
  62. Wu, R.F.; Ma, Z.; Myers, D.P.; Terada, L.S. HIV-1 Tat activates dual Nox pathways leading to independent activation of ERK and JNK MAP kinases. J. Biol. Chem. 2007, 282, 37412–37419. [Google Scholar] [CrossRef]
  63. Nacken, W.; Anhlan, D.; Hrincius, E.; Mostafa, A.; Wolff, T.; Sadewasser, A.; Pleschka, S.; Ehrhardt, C.; Ludwig, S. Activation of c-jun N-terminal kinase upon influenza A virus (IAV) infection is independent of pathogen-related receptors but dependent on amino acid sequence variations of IAV NS1. J. Virol. 2014, 88, 8843–8852. [Google Scholar] [CrossRef] [PubMed]
  64. Kyriakopoulos, A.; Nigh, G.; McCullough, P.; Seneff, S. Mitogen Activated Protein Kinase (MAPK) Activation, p53, and Autophagy Inhibition Characterize the Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2) Spike Protein Induced Neurotoxicity. Cureus 2022, 14, e32361. [Google Scholar] [CrossRef] [PubMed]
  65. Chen, Y.; Li, N.; Yang, J.; Li, K.; Tang, M.; Zhao, X.; Guo, W.; Tong, A.; Nie, C.; Peng, Y.; et al. PUMA overexpression dissociates thioredoxin from ASK1 to activate the JNK/BCL-2/BCL-XL pathway augmenting apoptosis in ovarian cancer. Biochim. Biophys. Acta Mol. Basis Dis. 2022, 1868, 166553. [Google Scholar] [CrossRef] [PubMed]
  66. Wang, Y.; Su, B.; Sah, V.; Brown, J.; Han, J.; Chien, K. Cardiac hypertrophy induced by mitogen-activated protein kinase kinase 7, a specific activator for c-Jun NH2-terminal kinase in ventricular muscle cells. J. Biol. Chem. 1998, 273, 5423–5426. [Google Scholar] [CrossRef] [PubMed]
  67. Tsaousidou, E.; Paeger, L.; Belgardt, B.; Pal, M.; Wunderlich, C.; Brönneke, H.; Collienne, U.; Hampel, B.; Wunderlich, F.; Schmidt-Supprian, M.; et al. Distinct Roles for JNK and IKK Activation in Agouti-Related Peptide Neurons in the Development of Obesity and Insulin Resistance. Cell Rep. 2014, 9, 1495–1506. [Google Scholar] [CrossRef] [PubMed]
  68. Yung, J.H.M.; Giacca, A. Role of c-Jun N-terminal Kinase (JNK) in Obesity and Type 2 Diabetes. Cells 2020, 9, 706. [Google Scholar] [CrossRef] [PubMed]
  69. Hill, J.O. Understanding and addressing the epidemic of obesity: An energy balance perspective. Endocr. Rev. 2006, 27, 750–761. [Google Scholar] [CrossRef]
  70. Jung, U.J.; Choi, M.S. Obesity and its metabolic complications: The role of adipokines and the relationship between obesity, inflammation, insulin resistance, dyslipidemia and nonalcoholic fatty liver disease. Int. J. Mol. Sci. 2014, 15, 6184–6223. [Google Scholar] [CrossRef]
  71. Pi-Sunyer, X. The medical risks of obesity. Postgrad. Med. 2009, 121, 21–33. [Google Scholar] [CrossRef]
  72. Hirosumi, J.; Tuncman, G.; Chang, L.; Gorgun, C.Z.; Uysal, K.T.; Maeda, K.; Karin, M.; Hotamisligil, G.S. A central role for JNK in obesity and insulin resistance. Nature 2002, 420, 333–336. [Google Scholar] [CrossRef]
  73. Musi, N.; Goodyear, L.J. Insulin resistance and improvements in signal transduction. Endocrine 2006, 29, 73–80. [Google Scholar] [CrossRef]
  74. Nguyen, M.T.; Satoh, H.; Favelyukis, S.; Babendure, J.L.; Imamura, T.; Sbodio, J.I.; Zalevsky, J.; Dahiyat, B.I.; Chi, N.W.; Olefsky, J.M. JNK and tumor necrosis factor-alpha mediate free fatty acid-induced insulin resistance in 3T3-L1 adipocytes. J. Biol. Chem. 2005, 280, 35361–35371. [Google Scholar] [CrossRef]
  75. Kant, S.; Barrett, T.; Vertii, A.; Noh, Y.H.; Jung, D.Y.; Kim, J.K.; Davis, R.J. Role of the mixed-lineage protein kinase pathway in the metabolic stress response to obesity. Cell Rep. 2013, 4, 681–688. [Google Scholar] [CrossRef]
  76. Holzer, R.G.; Park, E.J.; Li, N.; Tran, H.; Chen, M.; Choi, C.; Solinas, G.; Karin, M. Saturated fatty acids induce c-Src clustering within membrane subdomains, leading to JNK activation. Cell 2011, 147, 173–184. [Google Scholar] [CrossRef]
  77. Kant, S.; Standen, C.L.; Morel, C.; Jung, D.Y.; Kim, J.K.; Swat, W.; Flavell, R.A.; Davis, R.J. A Protein Scaffold Coordinates SRC-Mediated JNK Activation in Response to Metabolic Stress. Cell Rep. 2017, 20, 2775–2783. [Google Scholar] [CrossRef]
  78. Jaeschke, A.; Czech, M.P.; Davis, R.J. An essential role of the JIP1 scaffold protein for JNK activation in adipose tissue. Genes. Dev. 2004, 18, 1976–1980. [Google Scholar] [CrossRef] [PubMed]
  79. Menacho-Marquez, M.; Nogueiras, R.; Fabbiano, S.; Sauzeau, V.; Al-Massadi, O.; Dieguez, C.; Bustelo, X.R. Chronic sympathoexcitation through loss of Vav3, a Rac1 activator, results in divergent effects on metabolic syndrome and obesity depending on diet. Cell Metab. 2013, 18, 199–211. [Google Scholar] [CrossRef] [PubMed]
  80. Solinas, G.; Becattini, B. JNK at the crossroad of obesity, insulin resistance, and cell stress response. Mol. Metab. 2017, 6, 174–184. [Google Scholar] [CrossRef] [PubMed]
  81. Tuncman, G.; Hirosumi, J.; Solinas, G.; Chang, L.; Karin, M.; Hotamisligil, G.S. Functional in vivo interactions between JNK1 and JNK2 isoforms in obesity and insulin resistance. Proc. Natl. Acad. Sci. USA 2006, 103, 10741–10746. [Google Scholar] [CrossRef] [PubMed]
  82. Waeber, G.; Delplanque, J.; Bonny, C.; Mooser, V.; Steinmann, M.; Widmann, C.; Maillard, A.; Miklossy, J.; Dina, C.; Hani, E.H.; et al. The gene MAPK8IP1, encoding islet-brain-1, is a candidate for type 2 diabetes. Nat. Genet. 2000, 24, 291–295. [Google Scholar] [CrossRef] [PubMed]
  83. Hotamisligil, G.S.; Shargill, N.S.; Spiegelman, B.M. Adipose expression of tumor necrosis factor-alpha: Direct role in obesity-linked insulin resistance. Science 1993, 259, 87–91. [Google Scholar] [CrossRef] [PubMed]
  84. Solinas, G.; Karin, M. JNK1 and IKKbeta: Molecular links between obesity and metabolic dysfunction. FASEB J. 2010, 24, 2596–2611. [Google Scholar] [CrossRef]
  85. Tanti, J.F.; Gremeaux, T.; van Obberghen, E.; Le Marchand-Brustel, Y. Serine/threonine phosphorylation of insulin receptor substrate 1 modulates insulin receptor signaling. J. Biol. Chem. 1994, 269, 6051–6057. [Google Scholar] [CrossRef]
  86. Aguirre, V.; Uchida, T.; Yenush, L.; Davis, R.; White, M.F. The c-Jun NH(2)-terminal kinase promotes insulin resistance during association with insulin receptor substrate-1 and phosphorylation of Ser(307). J. Biol. Chem. 2000, 275, 9047–9054. [Google Scholar] [CrossRef]
  87. Aguirre, V.; Werner, E.D.; Giraud, J.; Lee, Y.H.; Shoelson, S.E.; White, M.F. Phosphorylation of Ser307 in insulin receptor substrate-1 blocks interactions with the insulin receptor and inhibits insulin action. J. Biol. Chem. 2002, 277, 1531–1537. [Google Scholar] [CrossRef]
  88. Solinas, G.; Naugler, W.; Galimi, F.; Lee, M.S.; Karin, M. Saturated fatty acids inhibit induction of insulin gene transcription by JNK-mediated phosphorylation of insulin-receptor substrates. Proc. Natl. Acad. Sci. USA 2006, 103, 16454–16459. [Google Scholar] [CrossRef]
  89. Sharfi, H.; Eldar-Finkelman, H. Sequential phosphorylation of insulin receptor substrate-2 by glycogen synthase kinase-3 and c-Jun NH2-terminal kinase plays a role in hepatic insulin signaling. Am. J. Physiol. Endocrinol. Metab. 2008, 294, E307–E315. [Google Scholar] [CrossRef]
  90. Han, M.S.; Jung, D.Y.; Morel, C.; Lakhani, S.A.; Kim, J.K.; Flavell, R.A.; Davis, R.J. JNK expression by macrophages promotes obesity-induced insulin resistance and inflammation. Science 2013, 339, 218–222. [Google Scholar] [CrossRef]
  91. Belgardt, B.F.; Mauer, J.; Wunderlich, F.T.; Ernst, M.B.; Pal, M.; Spohn, G.; Bronneke, H.S.; Brodesser, S.; Hampel, B.; Schauss, A.C.; et al. Hypothalamic and pituitary c-Jun N-terminal kinase 1 signaling coordinately regulates glucose metabolism. Proc. Natl. Acad. Sci. USA 2010, 107, 6028–6033. [Google Scholar] [CrossRef] [PubMed]
  92. Sabio, G.; Cavanagh-Kyros, J.; Barrett, T.; Jung, D.Y.; Ko, H.J.; Ong, H.; Morel, C.; Mora, A.; Reilly, J.; Kim, J.K.; et al. Role of the hypothalamic-pituitary-thyroid axis in metabolic regulation by JNK1. Genes. Dev. 2010, 24, 256–264. [Google Scholar] [CrossRef] [PubMed]
  93. Vernia, S.; Cavanagh-Kyros, J.; Barrett, T.; Jung, D.Y.; Kim, J.K.; Davis, R.J. Diet-induced obesity mediated by the JNK/DIO2 signal transduction pathway. Genes. Dev. 2013, 27, 2345–2355. [Google Scholar] [CrossRef] [PubMed]
  94. Vernia, S.; Morel, C.; Madara, J.C.; Cavanagh-Kyros, J.; Barrett, T.; Chase, K.; Kennedy, N.J.; Jung, D.Y.; Kim, J.K.; Aronin, N.; et al. Excitatory transmission onto AgRP neurons is regulated by cJun NH2-terminal kinase 3 in response to metabolic stress. Elife 2016, 5, e10031. [Google Scholar] [CrossRef] [PubMed]
  95. Vernia, S.; Cavanagh-Kyros, J.; Garcia-Haro, L.; Sabio, G.; Barrett, T.; Jung, D.Y.; Kim, J.K.; Xu, J.; Shulha, H.P.; Garber, M.; et al. The PPARalpha-FGF21 hormone axis contributes to metabolic regulation by the hepatic JNK signaling pathway. Cell Metab. 2014, 20, 512–525. [Google Scholar] [CrossRef] [PubMed]
  96. Vernia, S.; Cavanagh-Kyros, J.; Barrett, T.; Tournier, C.; Davis, R.J. Fibroblast Growth Factor 21 Mediates Glycemic Regulation by Hepatic JNK. Cell Rep. 2016, 14, 2273–2280. [Google Scholar] [CrossRef]
  97. Lanuza-Masdeu, J.; Arevalo, M.I.; Vila, C.; Barbera, A.; Gomis, R.; Caelles, C. In vivo JNK activation in pancreatic beta-cells leads to glucose intolerance caused by insulin resistance in pancreas. Diabetes 2013, 62, 2308–2317. [Google Scholar] [CrossRef]
  98. Hui, D.Y. A no-no for NonO and JNK in extracellular matrix homeostasis and vascular stability. Arterioscler. Thromb. Vasc. Biol. 2007, 27, 1677–1678. [Google Scholar] [CrossRef]
  99. Ricci, R.; Sumara, G.; Sumara, I.; Rozenberg, I.; Kurrer, M.; Akhmedov, A.; Hersberger, M.; Eriksson, U.; Eberli, F.R.; Becher, B.; et al. Requirement of JNK2 for scavenger receptor A-mediated foam cell formation in atherogenesis. Science 2004, 306, 1558–1561. [Google Scholar] [CrossRef]
  100. Amini, N.; Boyle, J.J.; Moers, B.; Warboys, C.M.; Malik, T.H.; Zakkar, M.; Francis, S.E.; Mason, J.C.; Haskard, D.O.; Evans, P.C. Requirement of JNK1 for endothelial cell injury in atherogenesis. Atherosclerosis 2014, 235, 613–618. [Google Scholar] [CrossRef]
  101. 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]
  102. Czaja, M.J. JNK regulation of hepatic manifestations of the metabolic syndrome. Trends Endocrinol. Metab. 2010, 21, 707–713. [Google Scholar] [CrossRef]
  103. Hirata, Y.; Inoue, A.; Suzuki, S.; Takahashi, M.; Matsui, R.; Kono, N.; Noguchi, T.; Matsuzawa, A. trans-Fatty acids facilitate DNA damage-induced apoptosis through the mitochondrial JNK-Sab-ROS positive feedback loop. Sci. Rep. 2020, 10, 2743. [Google Scholar] [CrossRef]
  104. Kluwe, J.; Pradere, J.P.; Gwak, G.Y.; Mencin, A.; De Minicis, S.; Osterreicher, C.H.; Colmenero, J.; Bataller, R.; Schwabe, R.F. Modulation of hepatic fibrosis by c-Jun-N-terminal kinase inhibition. Gastroenterology 2010, 138, 347–359. [Google Scholar] [CrossRef]
  105. Jiang, Y.; Xu, J.; Huang, P.; Yang, L.; Liu, Y.; Li, Y.; Wang, J.; Song, H.; Zheng, P. Scoparone Improves Nonalcoholic Steatohepatitis Through Alleviating JNK/Sab Signaling Pathway-Mediated Mitochondrial Dysfunction. Front. Pharmacol. 2022, 13, 863756. [Google Scholar] [CrossRef]
  106. Wei, Y.; Pagliassotti, M.J. Hepatospecific effects of fructose on c-jun NH2-terminal kinase: Implications for hepatic insulin resistance. Am. J. Physiol. Endocrinol. Metab. 2004, 287, E926–E933. [Google Scholar] [CrossRef] [PubMed]
  107. Win, S.; Than, T.A.; Zhang, J.; Oo, C.; Min, R.W.M.; Kaplowitz, N. New insights into the role and mechanism of c-Jun-N-terminal kinase signaling in the pathobiology of liver diseases. Hepatology 2018, 67, 2013–2024. [Google Scholar] [CrossRef] [PubMed]
  108. Wang, Q.; Zhou, H.; Bu, Q.; Wei, S.; Li, L.; Zhou, J.; Zhou, S.; Su, W.; Liu, M.; Liu, Z.; et al. Role of XBP1 in regulating the progression of non-alcoholic steatohepatitis. J. Hepatol. 2022, 77, 312–325. [Google Scholar] [CrossRef] [PubMed]
  109. Guo, L.; Guo, Y.Y.; Li, B.Y.; Peng, W.Q.; Chang, X.X.; Gao, X.; Tang, Q.Q. Enhanced acetylation of ATP-citrate lyase promotes the progression of nonalcoholic fatty liver disease. J. Biol. Chem. 2019, 294, 11805–11816. [Google Scholar] [CrossRef]
  110. Win, S.; Than, T.A.; Kaplowitz, N. The Regulation of JNK Signaling Pathways in Cell Death through the Interplay with Mitochondrial SAB and Upstream Post-Translational Effects. Int. J. Mol. Sci. 2018, 19, 3657. [Google Scholar] [CrossRef]
  111. Wang, P.X.; Ji, Y.X.; Zhang, X.J.; Zhao, L.P.; Yan, Z.Z.; Zhang, P.; Shen, L.J.; Yang, X.; Fang, J.; Tian, S.; et al. Targeting CASP8 and FADD-like apoptosis regulator ameliorates nonalcoholic steatohepatitis in mice and nonhuman primates. Nat. Med. 2017, 23, 439–449. [Google Scholar] [CrossRef] [PubMed]
  112. Ibrahim, S.H.; Hirsova, P.; Tomita, K.; Bronk, S.F.; Werneburg, N.W.; Harrison, S.A.; Goodfellow, V.S.; Malhi, H.; Gores, G.J. Mixed lineage kinase 3 mediates release of C-X-C motif ligand 10-bearing chemotactic extracellular vesicles from lipotoxic hepatocytes. Hepatology 2016, 63, 731–744. [Google Scholar] [CrossRef] [PubMed]
  113. Han, M.S.; Barrett, T.; Brehm, M.A.; Davis, R.J. Inflammation Mediated by JNK in Myeloid Cells Promotes the Development of Hepatitis and Hepatocellular Carcinoma. Cell Rep. 2016, 15, 19–26. [Google Scholar] [CrossRef] [PubMed]
  114. de Los Reyes Corrales, T.; Losada-Perez, M.; Casas-Tinto, S. JNK Pathway in CNS Pathologies. Int. J. Mol. Sci. 2021, 22, 3883. [Google Scholar] [CrossRef] [PubMed]
  115. La Marca, J.E.; Richardson, H.E. Two-Faced: Roles of JNK Signalling During Tumourigenesis in the Drosophila Model. Front. Cell Dev. Biol. 2020, 8, 42. [Google Scholar] [CrossRef] [PubMed]
  116. Musi, C.A.; Agro, G.; Santarella, F.; Iervasi, E.; Borsello, T. JNK3 as Therapeutic Target and Biomarker in Neurodegenerative and Neurodevelopmental Brain Diseases. Cells 2020, 9, 2190. [Google Scholar] [CrossRef] [PubMed]
  117. Portela, M.; Venkataramani, V.; Fahey-Lozano, N.; Seco, E.; Losada-Perez, M.; Winkler, F.; Casas-Tinto, S. Glioblastoma cells vampirize WNT from neurons and trigger a JNK/MMP signaling loop that enhances glioblastoma progression and neurodegeneration. PLoS Biol. 2019, 17, e3000545. [Google Scholar] [CrossRef] [PubMed]
  118. Gallego, O. Nonsurgical treatment of recurrent glioblastoma. Curr. Oncol. 2015, 22, e273–e281. [Google Scholar] [CrossRef]
  119. Kitanaka, C.; Sato, A.; Okada, M. JNK Signaling in the Control of the Tumor-Initiating Capacity Associated with Cancer Stem Cells. Genes. Cancer 2013, 4, 388–396. [Google Scholar] [CrossRef]
  120. Matsuda, K.; Sato, A.; Okada, M.; Shibuya, K.; Seino, S.; Suzuki, K.; Watanabe, E.; Narita, Y.; Shibui, S.; Kayama, T.; et al. Targeting JNK for therapeutic depletion of stem-like glioblastoma cells. Sci. Rep. 2012, 2, 516. [Google Scholar] [CrossRef] [PubMed]
  121. Portela, M.; Mitchell, T.; Casas-Tinto, S. Cell-to-cell communication mediates glioblastoma progression in Drosophila. Biol. Open 2020, 9, bio053405. [Google Scholar] [CrossRef]
  122. Jarabo, P.; de Pablo, C.; Herranz, H.; Martin, F.A.; Casas-Tinto, S. Insulin signaling mediates neurodegeneration in glioma. Life Sci. Alliance 2021, 4, e202000693. [Google Scholar] [CrossRef]
  123. Hepp Rehfeldt, S.C.; Majolo, F.; Goettert, M.I.; Laufer, S. c-Jun N-Terminal Kinase Inhibitors as Potential Leads for New Therapeutics for Alzheimer’s Diseases. Int. J. Mol. Sci. 2020, 21, 9677. [Google Scholar] [CrossRef] [PubMed]
  124. Yao, M.; Nguyen, T.V.; Pike, C.J. Beta-amyloid-induced neuronal apoptosis involves c-Jun N-terminal kinase-dependent downregulation of Bcl-w. J. Neurosci. 2005, 25, 1149–1158. [Google Scholar] [CrossRef] [PubMed]
  125. Yoshida, H.; Hastie, C.J.; McLauchlan, H.; Cohen, P.; Goedert, M. Phosphorylation of microtubule-associated protein tau by isoforms of c-Jun N-terminal kinase (JNK). J. Neurochem. 2004, 90, 352–358. [Google Scholar] [CrossRef] [PubMed]
  126. Bozyczko-Coyne, D.; Saporito, M.S.; Hudkins, R.L. Targeting the JNK pathway for therapeutic benefit in CNS disease. Curr. Drug Targets CNS Neurol. Disord. 2002, 1, 31–49. [Google Scholar] [CrossRef] [PubMed]
  127. Morishima, Y.; Gotoh, Y.; Zieg, J.; Barrett, T.; Takano, H.; Flavell, R.; Davis, R.J.; Shirasaki, Y.; Greenberg, M.E. Beta-amyloid induces neuronal apoptosis via a mechanism that involves the c-Jun N-terminal kinase pathway and the induction of Fas ligand. J. Neurosci. 2001, 21, 7551–7560. [Google Scholar] [CrossRef]
  128. Choi, W.S.; Abel, G.; Klintworth, H.; Flavell, R.A.; Xia, Z. JNK3 mediates paraquat- and rotenone-induced dopaminergic neuron death. J. Neuropathol. Exp. Neurol. 2010, 69, 511–520. [Google Scholar] [CrossRef]
  129. Bekker, M.; Abrahams, S.; Loos, B.; Bardien, S. Can the interplay between autophagy and apoptosis be targeted as a novel therapy for Parkinson’s disease? Neurobiol. Aging 2021, 100, 91–105. [Google Scholar] [CrossRef]
  130. Sieradzan, K.A.; Mann, D.M. The selective vulnerability of nerve cells in Huntington’s disease. Neuropathol. Appl. Neurobiol. 2001, 27, 1–21. [Google Scholar] [CrossRef]
  131. Morfini, G.A.; You, Y.M.; Pollema, S.L.; Kaminska, A.; Liu, K.; Yoshioka, K.; Bjorkblom, B.; Coffey, E.T.; Bagnato, C.; Han, D.; et al. Pathogenic huntingtin inhibits fast axonal transport by activating JNK3 and phosphorylating kinesin. Nat. Neurosci. 2009, 12, 864–871. [Google Scholar] [CrossRef]
  132. Perrin, V.; Dufour, N.; Raoul, C.; Hassig, R.; Brouillet, E.; Aebischer, P.; Luthi-Carter, R.; Deglon, N. Implication of the JNK pathway in a rat model of Huntington’s disease. Exp. Neurol. 2009, 215, 191–200. [Google Scholar] [CrossRef] [PubMed]
  133. Yang, D.D.; Kuan, C.Y.; Whitmarsh, A.J.; Rincon, M.; Zheng, T.S.; Davis, R.J.; Rakic, P.; Flavell, R.A. Absence of excitotoxicity-induced apoptosis in the hippocampus of mice lacking the Jnk3 gene. Nature 1997, 389, 865–870. [Google Scholar] [CrossRef] [PubMed]
  134. Zhang, P.; Miller, B.S.; Rosenzweig, S.A.; Bhat, N.R. Activation of C-jun N-terminal kinase/stress-activated protein kinase in primary glial cultures. J. Neurosci. Res. 1996, 46, 114–121. [Google Scholar] [CrossRef]
  135. Bagnoud, M.; Briner, M.; Remlinger, J.; Meli, I.; Schuetz, S.; Pistor, M.; Salmen, A.; Chan, A.; Hoepner, R. c-Jun N-Terminal Kinase as a Therapeutic Target in Experimental Autoimmune Encephalomyelitis. Cells 2020, 9, 2154. [Google Scholar] [CrossRef] [PubMed]
  136. Guo, X.; Harada, C.; Namekata, K.; Matsuzawa, A.; Camps, M.; Ji, H.; Swinnen, D.; Jorand-Lebrun, C.; Muzerelle, M.; Vitte, P.; et al. Regulation of the severity of neuroinflammation and demyelination by TLR-ASK1-p38 pathway. EMBO Mol. Med. 2010, 2, 504–515. [Google Scholar] [CrossRef]
  137. Singh, A.; Upadhayay, S.; Mehan, S. Understanding Abnormal c-JNK/p38MAPK Signaling Overactivation Involved in the Progression of Multiple Sclerosis: Possible Therapeutic Targets and Impact on Neurodegenerative Diseases. Neurotox. Res. 2021, 39, 1630–1650. [Google Scholar] [CrossRef]
  138. Lindwall, C.; Kanje, M. Retrograde axonal transport of JNK signaling molecules influence injury induced nuclear changes in p-c-Jun and ATF3 in adult rat sensory neurons. Mol. Cell Neurosci. 2005, 29, 269–282. [Google Scholar] [CrossRef]
  139. Pathak, A.; Clark, S.; Bronfman, F.C.; Deppmann, C.D.; Carter, B.D. Long-distance regressive signaling in neural development and disease. Wiley Interdiscip. Rev. Dev. Biol. 2021, 10, e382. [Google Scholar] [CrossRef] [PubMed]
  140. Macdonald, J.M.; Doherty, J.; Hackett, R.; Freeman, M.R. The c-Jun kinase signaling cascade promotes glial engulfment activity through activation of draper and phagocytic function. Cell Death Differ. 2013, 20, 1140–1148. [Google Scholar] [CrossRef]
  141. Huntwork-Rodriguez, S.; Wang, B.; Watkins, T.; Ghosh, A.S.; Pozniak, C.D.; Bustos, D.; Newton, K.; Kirkpatrick, D.S.; Lewcock, J.W. JNK-mediated phosphorylation of DLK suppresses its ubiquitination to promote neuronal apoptosis. J. Cell Biol. 2013, 202, 747–763. [Google Scholar] [CrossRef]
  142. Valakh, V.; Frey, E.; Babetto, E.; Walker, L.J.; DiAntonio, A. Cytoskeletal disruption activates the DLK/JNK pathway, which promotes axonal regeneration and mimics a preconditioning injury. Neurobiol. Dis. 2015, 77, 13–25. [Google Scholar] [CrossRef] [PubMed]
  143. Hirai, S.; Banba, Y.; Satake, T.; Ohno, S. Axon formation in neocortical neurons depends on stage-specific regulation of microtubule stability by the dual leucine zipper kinase-c-Jun N-terminal kinase pathway. J. Neurosci. 2011, 31, 6468–6480. [Google Scholar] [CrossRef] [PubMed]
  144. Hellal, F.; Hurtado, A.; Ruschel, J.; Flynn, K.C.; Laskowski, C.J.; Umlauf, M.; Kapitein, L.C.; Strikis, D.; Lemmon, V.; Bixby, J.; et al. Microtubule stabilization reduces scarring and causes axon regeneration after spinal cord injury. Science 2011, 331, 928–931. [Google Scholar] [CrossRef]
  145. Apara, A.; Galvao, J.; Wang, Y.; Blackmore, M.; Trillo, A.; Iwao, K.; Brown, D.P., Jr.; Fernandes, K.A.; Huang, A.; Nguyen, T.; et al. KLF9 and JNK3 Interact to Suppress Axon Regeneration in the Adult CNS. J. Neurosci. 2017, 37, 9632–9644. [Google Scholar] [CrossRef]
  146. Lai, B.; Wu, C.H.; Lai, J.H. Activation of c-Jun N-Terminal Kinase, a Potential Therapeutic Target in Autoimmune Arthritis. Cells 2020, 9, 2466. [Google Scholar] [CrossRef] [PubMed]
  147. Smolen, J.S.; Aletaha, D.; Barton, A.; Burmester, G.R.; Emery, P.; Firestein, G.S.; Kavanaugh, A.; McInnes, I.B.; Solomon, D.H.; Strand, V.; et al. Rheumatoid arthritis. Nat. Rev. Dis. Primers 2018, 4, 18001. [Google Scholar] [CrossRef] [PubMed]
  148. Fukushima, A.; Boyle, D.L.; Corr, M.; Firestein, G.S. Kinetic analysis of synovial signalling and gene expression in animal models of arthritis. Ann. Rheum. Dis. 2010, 69, 918–923. [Google Scholar] [CrossRef]
  149. Han, Z.; Boyle, D.L.; Chang, L.; Bennett, B.; Karin, M.; Yang, L.; Manning, A.M.; Firestein, G.S. c-Jun N-terminal kinase is required for metalloproteinase expression and joint destruction in inflammatory arthritis. J. Clin. Invest. 2001, 108, 73–81. [Google Scholar] [CrossRef]
  150. Schett, G.; Tohidast-Akrad, M.; Smolen, J.S.; Schmid, B.J.; Steiner, C.W.; Bitzan, P.; Zenz, P.; Redlich, K.; Xu, Q.; Steiner, G. Activation, differential localization, and regulation of the stress-activated protein kinases, extracellular signal-regulated kinase, c-JUN N-terminal kinase, and p38 mitogen-activated protein kinase, in synovial tissue and cells in rheumatoid arthritis. Arthritis Rheum. 2000, 43, 2501–2512. [Google Scholar] [CrossRef]
  151. Lories, R.J.; Derese, I.; Luyten, F.P.; de Vlam, K. Activation of nuclear factor kappa B and mitogen activated protein kinases in psoriatic arthritis before and after etanercept treatment. Clin. Exp. Rheumatol. 2008, 26, 96–102. [Google Scholar]
  152. Li, X.; Wang, J.; Zhan, Z.; Li, S.; Zheng, Z.; Wang, T.; Zhang, K.; Pan, H.; Li, Z.; Zhang, N.; et al. Inflammation Intensity-Dependent Expression of Osteoinductive Wnt Proteins Is Critical for Ectopic New Bone Formation in Ankylosing Spondylitis. Arthritis Rheumatol. 2018, 70, 1056–1070. [Google Scholar] [CrossRef]
  153. Ge, H.X.; Zou, F.M.; Li, Y.; Liu, A.M.; Tu, M. JNK pathway in osteoarthritis: Pathological and therapeutic aspects. J. Recept. Signal Transduct. Res. 2017, 37, 431–436. [Google Scholar] [CrossRef]
  154. Im, H.J.; Muddasani, P.; Natarajan, V.; Schmid, T.M.; Block, J.A.; Davis, F.; van Wijnen, A.J.; Loeser, R.F. Basic fibroblast growth factor stimulates matrix metalloproteinase-13 via the molecular cross-talk between the mitogen-activated protein kinases and protein kinase Cdelta pathways in human adult articular chondrocytes. J. Biol. Chem. 2007, 282, 11110–11121. [Google Scholar] [CrossRef]
  155. Yang, P.; Tan, J.; Yuan, Z.; Meng, G.; Bi, L.; Liu, J. Expression profile of cytokines and chemokines in osteoarthritis patients: Proinflammatory roles for CXCL8 and CXCL11 to chondrocytes. Int. Immunopharmacol. 2016, 40, 16–23. [Google Scholar] [CrossRef]
  156. Roy, P.K.; Rashid, F.; Bragg, J.; Ibdah, J.A. Role of the JNK signal transduction pathway in inflammatory bowel disease. World J. Gastroenterol. 2008, 14, 200–202. [Google Scholar] [CrossRef]
  157. Mitsuyama, K.; Suzuki, A.; Tomiyasu, N.; Tsuruta, O.; Kitazaki, S.; Takeda, T.; Satoh, Y.; Bennett, B.L.; Toyonaga, A.; Sata, M. Pro-inflammatory signaling by Jun-N-terminal kinase in inflammatory bowel disease. Int. J. Mol. Med. 2006, 17, 449–455. [Google Scholar] [CrossRef] [PubMed]
  158. Chromik, A.M.; Muller, A.M.; Korner, J.; Belyaev, O.; Holland-Letz, T.; Schmitz, F.; Herdegen, T.; Uhl, W.; Mittelkotter, U. Genetic deletion of JNK1 and JNK2 aggravates the DSS-induced colitis in mice. J. Invest. Surg. 2007, 20, 23–33. [Google Scholar] [CrossRef] [PubMed]
  159. Zhi, Y.; Zhou, X.; Yu, J.; Yuan, L.; Zhang, H.; Ng, D.C.H.; Xu, Z.; Xu, D. Pathophysiological Significance of WDR62 and JNK Signaling in Human Diseases. Front. Cell Dev. Biol. 2021, 9, 640753. [Google Scholar] [CrossRef] [PubMed]
  160. Hideshima, T.; Hayashi, T.; Chauhan, D.; Akiyama, M.; Richardson, P.; Anderson, K. Biologic sequelae of c-Jun NH(2)-terminal kinase (JNK) activation in multiple myeloma cell lines. Oncogene 2003, 22, 8797–8801. [Google Scholar] [CrossRef]
  161. Sharkey, J.; Khong, T.; Spencer, A. PKC412 demonstrates JNK-dependent activity against human multiple myeloma cells. Blood 2007, 109, 1712–1719. [Google Scholar] [CrossRef]
  162. Sung, Y.; Park, S.; Park, S.J.; Jeong, J.; Choi, M.; Lee, J.; Kwon, W.; Jang, S.; Lee, M.H.; Kim, D.J.; et al. Jazf1 promotes prostate cancer progression by activating JNK/Slug. Oncotarget 2018, 9, 755–765. [Google Scholar] [CrossRef]
  163. Du, R.; Tang, G.; Tang, Z.; Kuang, Y. Ectopic expression of CC chemokine receptor 7 promotes prostate cancer cells metastasis via Notch1 signaling. J. Cell Biochem. 2019, 120, 9639–9647. [Google Scholar] [CrossRef] [PubMed]
  164. Gkouveris, I.; Nikitakis, N.G. Role of JNK signaling in oral cancer: A mini review. Tumour Biol. 2017, 39, 1010428317711659. [Google Scholar] [CrossRef] [PubMed]
  165. Noutomi, T.; Itoh, M.; Toyota, H.; Takada, E.; Mizuguchi, J. Tumor necrosis factor-related apoptosis-inducing ligand induces apoptotic cell death through c-Jun NH2-terminal kinase activation in squamous cell carcinoma cells. Oncol. Rep. 2009, 22, 1169–1172. [Google Scholar] [CrossRef] [PubMed]
  166. Kim, H.J.; Chakravarti, N.; Oridate, N.; Choe, C.; Claret, F.X.; Lotan, R. N-(4-hydroxyphenyl)retinamide-induced apoptosis triggered by reactive oxygen species is mediated by activation of MAPKs in head and neck squamous carcinoma cells. Oncogene 2006, 25, 2785–2794. [Google Scholar] [CrossRef] [PubMed]
  167. Gross, N.D.; Boyle, J.O.; Du, B.; Kekatpure, V.D.; Lantowski, A.; Thaler, H.T.; Weksler, B.B.; Subbaramaiah, K.; Dannenberg, A.J. Inhibition of Jun NH2-terminal kinases suppresses the growth of experimental head and neck squamous cell carcinoma. Clin. Cancer Res. 2007, 13, 5910–5917. [Google Scholar] [CrossRef]
  168. Luo, Z.; Han, Z.; Shou, F.; Li, Y.; Chen, Y. LINC00958 Accelerates Cell Proliferation and Migration in Non-Small Cell Lung Cancer Through JNK/c-JUN Signaling. Hum. Gene Ther. Methods 2019, 30, 226–234. [Google Scholar] [CrossRef]
  169. Lin, X.; Ye, R.; Li, Z.; Zhang, B.; Huang, Y.; Du, J.; Wang, B.; Meng, H.; Xian, H.; Yang, X.; et al. KIAA1429 promotes tumorigenesis and gefitinib resistance in lung adenocarcinoma by activating the JNK/ MAPK pathway in an m(6)A-dependent manner. Drug Resist. Updat. 2023, 66, 100908. [Google Scholar] [CrossRef]
  170. Cellurale, C.; Girnius, N.; Jiang, F.; Cavanagh-Kyros, J.; Lu, S.; Garlick, D.S.; Mercurio, A.M.; Davis, R.J. Role of JNK in mammary gland development and breast cancer. Cancer Res. 2012, 72, 472–481. [Google Scholar] [CrossRef]
  171. Shen, K.; Xie, J.; Wang, H.; Zhang, H.; Yu, M.; Lu, F.; Tan, H.; Xu, H. Cambogin Induces Caspase-Independent Apoptosis through the ROS/JNK Pathway and Epigenetic Regulation in Breast Cancer Cells. Mol. Cancer Ther. 2015, 14, 1738–1749. [Google Scholar] [CrossRef]
  172. Itah, Z.; Chaudhry, S.; Raju Ponny, S.; Aydemir, O.; Lee, A.; Cavanagh-Kyros, J.; Tournier, C.; Muller, W.J.; Davis, R.J. HER2-driven breast cancer suppression by the JNK signaling pathway. Proc. Natl. Acad. Sci. USA 2023, 120, e2218373120. [Google Scholar] [CrossRef]
  173. Tong, C.; Yin, Z.; Song, Z.; Dockendorff, A.; Huang, C.; Mariadason, J.; Flavell, R.A.; Davis, R.J.; Augenlicht, L.H.; Yang, W. c-Jun NH2-terminal kinase 1 plays a critical role in intestinal homeostasis and tumor suppression. Am. J. Pathol. 2007, 171, 297–303. [Google Scholar] [CrossRef] [PubMed]
  174. Kwak, A.W.; Park, J.W.; Lee, S.O.; Lee, J.Y.; Seo, J.H.; Yoon, G.; Lee, M.H.; Choi, J.S.; Shim, J.H. Isolinderalactone sensitizes oxaliplatin-resistance colorectal cancer cells through JNK/p38 MAPK signaling pathways. Phytomedicine 2022, 105, 154383. [Google Scholar] [CrossRef] [PubMed]
  175. Choi, H.S.; Bode, A.M.; Shim, J.H.; Lee, S.Y.; Dong, Z. c-Jun N-terminal kinase 1 phosphorylates Myt1 to prevent UVA-induced skin cancer. Mol. Cell Biol. 2009, 29, 2168–2180. [Google Scholar] [CrossRef] [PubMed]
  176. Chen, N.; Nomura, M.; She, Q.B.; Ma, W.Y.; Bode, A.M.; Wang, L.; Flavell, R.A.; Dong, Z. Suppression of skin tumorigenesis in c-Jun NH(2)-terminal kinase-2-deficient mice. Cancer Res. 2001, 61, 3908–3912. [Google Scholar]
  177. Ding, X.; Wang, X.; Zhu, X.; Zhang, J.; Zhu, Y.; Shao, X.; Zhou, X. JNK/AP1 Pathway Regulates MYC Expression and BCR Signaling through Ig Enhancers in Burkitt Lymphoma Cells. J. Cancer 2020, 11, 610–618. [Google Scholar] [CrossRef] [PubMed]
  178. Yang, X.; Xiang, X.; Xia, M.; Su, J.; Wu, Y.; Shen, L.; Xu, Y.; Sun, L. Inhibition of JNK3 promotes apoptosis induced by BH3 mimetic S1 in chemoresistant human ovarian cancer cells. Anat Rec (Hoboken) 2015, 298, 386–395. [Google Scholar] [CrossRef] [PubMed]
  179. Vivas-Mejia, P.; Benito, J.M.; Fernandez, A.; Han, H.D.; Mangala, L.; Rodriguez-Aguayo, C.; Chavez-Reyes, A.; Lin, Y.G.; Carey, M.S.; Nick, A.M.; et al. c-Jun-NH2-kinase-1 inhibition leads to antitumor activity in ovarian cancer. Clin. Cancer Res. 2010, 16, 184–194. [Google Scholar] [CrossRef] [PubMed]
  180. Wei, L.; Zhu, S.; Ruan, G.; Hou, L.; Wang, J.; Wang, B.; Liu, J. Infectious bursal disease virus-induced activation of JNK signaling pathway is required for virus replication and correlates with virus-induced apoptosis. Virology 2011, 420, 156–163. [Google Scholar] [CrossRef]
  181. Ceballos-Olvera, I.; Chavez-Salinas, S.; Medina, F.; Ludert, J.E.; del Angel, R.M. JNK phosphorylation, induced during dengue virus infection, is important for viral infection and requires the presence of cholesterol. Virology 2010, 396, 30–36. [Google Scholar] [CrossRef] [PubMed]
  182. Zhang, J.; Ruan, T.; Sheng, T.; Wang, J.; Sun, J.; Wang, J.; Prinz, R.A.; Peng, D.; Liu, X.; Xu, X. Role of c-Jun terminal kinase (JNK) activation in influenza A virus-induced autophagy and replication. Virology 2019, 526, 1–12. [Google Scholar] [CrossRef]
  183. Chang, L.; Karin, M. Mammalian MAP kinase signalling cascades. Nature 2001, 410, 37–40. [Google Scholar] [CrossRef]
  184. Fu, J.; Peng, H. Lipopolysaccharides attenuates growth of HS cells through the JNK pathway. Cytotechnology 2016, 68, 2389–2394. [Google Scholar] [CrossRef]
  185. Nguyen, C.T.; Kim, E.H.; Luong, T.T.; Pyo, S.; Rhee, D.K. TLR4 mediates pneumolysin-induced ATF3 expression through the JNK/p38 pathway in Streptococcus pneumoniae-infected RAW 264.7 cells. Mol. Cells 2015, 38, 58–64. [Google Scholar] [CrossRef] [PubMed]
  186. Smith, W.E.; Kane, A.V.; Campbell, S.T.; Acheson, D.W.; Cochran, B.H.; Thorpe, C.M. Shiga toxin 1 triggers a ribotoxic stress response leading to p38 and JNK activation and induction of apoptosis in intestinal epithelial cells. Infect. Immun. 2003, 71, 1497–1504. [Google Scholar] [CrossRef] [PubMed]
  187. Jones, R.M.; Wu, H.; Wentworth, C.; Luo, L.; Collier-Hyams, L.; Neish, A.S. Salmonella AvrA Coordinates Suppression of Host Immune and Apoptotic Defenses via JNK Pathway Blockade. Cell Host Microbe 2008, 3, 233–244. [Google Scholar] [CrossRef]
  188. Geissler, A.; Haun, F.; Frank, D.O.; Wieland, K.; Simon, M.M.; Idzko, M.; Davis, R.J.; Maurer, U.; Borner, C. Apoptosis induced by the fungal pathogen gliotoxin requires a triple phosphorylation of Bim by JNK. Cell Death Differ. 2013, 20, 1317–1329. [Google Scholar] [CrossRef]
  189. Zhao, X.; Guo, Y.; Jiang, C.; Chang, Q.; Zhang, S.; Luo, T.; Zhang, B.; Jia, X.; Hung, M.C.; Dong, C.; et al. JNK1 negatively controls antifungal innate immunity by suppressing CD23 expression. Nat. Med. 2017, 23, 337–346. [Google Scholar] [CrossRef]
  190. Valere, A.; Garnotel, R.; Villena, I.; Guenounou, M.; Pinon, J.M.; Aubert, D. Activation of the cellular mitogen-activated protein kinase pathways ERK, P38 and JNK during Toxoplasma gondii invasion. Parasite 2003, 10, 59–64. [Google Scholar] [CrossRef]
  191. Sukhumavasi, W.; Warren, A.L.; Del Rio, L.; Denkers, E.Y. Absence of mitogen-activated protein kinase family member c-Jun N-terminal kinase-2 enhances resistance to Toxoplasma gondii. Exp. Parasitol. 2010, 126, 415–420. [Google Scholar] [CrossRef]
  192. Eshraghi, A.A.; Van de Water, T.R. Cochlear implantation trauma and noise-induced hearing loss: Apoptosis and therapeutic strategies. Anat. Rec. A Discov. Mol. Cell Evol. Biol. 2006, 288, 473–481. [Google Scholar] [CrossRef]
  193. Yamaguchi, O.; Higuchi, Y.; Hirotani, S.; Kashiwase, K.; Nakayama, H.; Hikoso, S.; Takeda, T.; Watanabe, T.; Asahi, M.; Taniike, M.; et al. Targeted deletion of apoptosis signal-regulating kinase 1 attenuates left ventricular remodeling. Proc. Natl. Acad. Sci. USA 2003, 100, 15883–15888. [Google Scholar] [CrossRef] [PubMed]
  194. Izumiya, Y.; Kim, S.; Izumi, Y.; Yoshida, K.; Yoshiyama, M.; Matsuzawa, A.; Ichijo, H.; Iwao, H. Apoptosis signal-regulating kinase 1 plays a pivotal role in angiotensin II-induced cardiac hypertrophy and remodeling. Circ. Res. 2003, 93, 874–883. [Google Scholar] [CrossRef] [PubMed]
  195. Dickhout, J.G.; Carlisle, R.E.; Austin, R.C. Interrelationship between cardiac hypertrophy, heart failure, and chronic kidney disease: Endoplasmic reticulum stress as a mediator of pathogenesis. Circ. Res. 2011, 108, 629–642. [Google Scholar] [CrossRef] [PubMed]
  196. Yoshimura, K.; Aoki, H.; Ikeda, Y.; Fujii, K.; Akiyama, N.; Furutani, A.; Hoshii, Y.; Tanaka, N.; Ricci, R.; Ishihara, T.; et al. Regression of abdominal aortic aneurysm by inhibition of c-Jun N-terminal kinase. Nat. Med. 2005, 11, 1330–1338. [Google Scholar] [CrossRef] [PubMed]
  197. Grynberg, K.; Ma, F.Y.; Nikolic-Paterson, D.J. The JNK Signaling Pathway in Renal Fibrosis. Front. Physiol. 2017, 8, 829. [Google Scholar] [CrossRef]
  198. Smith, A.O.; Jonassen, J.A.; Preval, K.M.; Davis, R.J.; Pazour, G.J. c-JUN n-Terminal Kinase (JNK) Signaling in Autosomal Dominant Polycystic Kidney Disease. J. Cell Signal 2022, 3, 62–78. [Google Scholar] [CrossRef]
  199. Stebbins, J.L.; De, S.K.; Machleidt, T.; Becattini, B.; Vazquez, J.; Kuntzen, C.; Chen, L.H.; Cellitti, J.F.; Riel-Mehan, M.; Emdadi, A.; et al. Identification of a new JNK inhibitor targeting the JNK-JIP interaction site. Proc. Natl. Acad. Sci. USA 2008, 105, 16809–16813. [Google Scholar] [CrossRef]
  200. Deloche, C.; Lopez-Lazaro, L.; Mouz, S.; Perino, J.; Abadie, C.; Combette, J.M. XG-102 administered to healthy male volunteers as a single intravenous infusion: A randomized, double-blind, placebo-controlled, dose-escalating study. Pharmacol. Res. Perspect. 2014, 2, e00020. [Google Scholar] [CrossRef] [PubMed]
  201. Li, C.; Miao, X.; Wang, S.; Adhikari, B.K.; Wang, X.; Sun, J.; Liu, Q.; Tong, Q.; Wang, Y. Novel Curcumin C66 That Protects Diabetes-Induced Aortic Damage Was Associated with Suppressing JNK2 and Upregulating Nrf2 Expression and Function. Oxid. Med. Cell Longev. 2018, 2018, 5783239. [Google Scholar] [CrossRef] [PubMed]
  202. Pan, Y.; Wang, Y.; Zhao, Y.; Peng, K.; Li, W.; Wang, Y.; Zhang, J.; Zhou, S.; Liu, Q.; Li, X.; et al. Inhibition of JNK phosphorylation by a novel curcumin analog prevents high glucose-induced inflammation and apoptosis in cardiomyocytes and the development of diabetic cardiomyopathy. Diabetes 2014, 63, 3497–3511. [Google Scholar] [CrossRef] [PubMed]
  203. Mohammadian Haftcheshmeh, S.; Karimzadeh, M.R.; Azhdari, S.; Vahedi, P.; Abdollahi, E.; Momtazi-Borojeni, A.A. Modulatory effects of curcumin on the atherogenic activities of inflammatory monocytes: Evidence from in vitro and animal models of human atherosclerosis. Biofactors 2020, 46, 341–355. [Google Scholar] [CrossRef] [PubMed]
  204. Cui, J.; Zhang, M.; Zhang, Y.Q.; Xu, Z.H. JNK pathway: Diseases and therapeutic potential. Acta Pharmacol. Sin. 2007, 28, 601–608. [Google Scholar] [CrossRef] [PubMed]
  205. Falsig, J.; Porzgen, P.; Lotharius, J.; Leist, M. Specific modulation of astrocyte inflammation by inhibition of mixed lineage kinases with CEP-1347. J. Immunol. 2004, 173, 2762–2770. [Google Scholar] [CrossRef] [PubMed]
  206. Lou, L.; Hu, D.; Chen, S.; Wang, S.; Xu, Y.; Huang, Y.; Shi, Y.; Zhang, H. Protective role of JNK inhibitor SP600125 in sepsis-induced acute lung injury. Int. J. Clin. Exp. Pathol. 2019, 12, 528–538. [Google Scholar] [PubMed]
  207. Kumar, A.; Singh, U.; Kini, S.; Garg, V.; Agrawal, S.; Tomar, P.; Pathak, P.; Chaudhary, A.; Gupta, P.; Malik, A. JNK pathway signaling: A novel and smarter therapeutic targets for various biological diseases. Future Med. Chem. 2015, 7, 2065–2086. [Google Scholar] [CrossRef]
  208. Mixed lineage kinase inhibitor CEP-1347 fails to delay disability in early Parkinson disease. Neurology 2007, 69, 1480–1490. [CrossRef]
  209. Gaillard, P.; Jeanclaude-Etter, I.; Ardissone, V.; Arkinstall, S.; Cambet, Y.; Camps, M.; Chabert, C.; Church, D.; Cirillo, R.; Gretener, D.; et al. Design and synthesis of the first generation of novel potent, selective, and in vivo active (benzothiazol-2-yl)acetonitrile inhibitors of the c-Jun N-terminal kinase. J. Med. Chem. 2005, 48, 4596–4607. [Google Scholar] [CrossRef]
  210. De, S.K.; Chen, L.H.; Stebbins, J.L.; Machleidt, T.; Riel-Mehan, M.; Dahl, R.; Chen, V.; Yuan, H.; Barile, E.; Emdadi, A.; et al. Discovery of 2-(5-nitrothiazol-2-ylthio)benzo[d]thiazoles as novel c-Jun N-terminal kinase inhibitors. Bioorg Med. Chem. 2009, 17, 2712–2717. [Google Scholar] [CrossRef]
  211. De, S.K.; Stebbins, J.L.; Chen, L.H.; Riel-Mehan, M.; Machleidt, T.; Dahl, R.; Yuan, H.; Emdadi, A.; Barile, E.; Chen, V.; et al. Design, synthesis, and structure-activity relationship of substrate competitive, selective, and in vivo active triazole and thiadiazole inhibitors of the c-Jun N-terminal kinase. J. Med. Chem. 2009, 52, 1943–1952. [Google Scholar] [CrossRef] [PubMed]
  212. Repici, M.; Mare, L.; Colombo, A.; Ploia, C.; Sclip, A.; Bonny, C.; Nicod, P.; Salmona, M.; Borsello, T. c-Jun N-terminal kinase binding domain-dependent phosphorylation of mitogen-activated protein kinase kinase 4 and mitogen-activated protein kinase kinase 7 and balancing cross-talk between c-Jun N-terminal kinase and extracellular signal-regulated kinase pathways in cortical neurons. Neuroscience 2009, 159, 94–103. [Google Scholar] [CrossRef]
  213. Kaneto, H.; Nakatani, Y.; Miyatsuka, T.; Kawamori, D.; Matsuoka, T.A.; Matsuhisa, M.; Kajimoto, Y.; Ichijo, H.; Yamasaki, Y.; Hori, M. Possible novel therapy for diabetes with cell-permeable JNK-inhibitory peptide. Nat. Med. 2004, 10, 1128–1132. [Google Scholar] [CrossRef] [PubMed]
  214. Bonny, C.; Oberson, A.; Negri, S.; Sauser, C.; Schorderet, D.F. Cell-permeable peptide inhibitors of JNK: Novel blockers of beta-cell death. Diabetes 2001, 50, 77–82. [Google Scholar] [CrossRef]
  215. Yarza, R.; Vela, S.; Solas, M.; Ramirez, M.J. c-Jun N-terminal Kinase (JNK) Signaling as a Therapeutic Target for Alzheimer’s Disease. Front. Pharmacol. 2015, 6, 321. [Google Scholar] [CrossRef] [PubMed]
  216. Stebbins, J.L.; De, S.K.; Pavlickova, P.; Chen, V.; Machleidt, T.; Chen, L.H.; Kuntzen, C.; Kitada, S.; Karin, M.; Pellecchia, M. Design and characterization of a potent and selective dual ATP- and substrate-competitive subnanomolar bidentate c-Jun N-terminal kinase (JNK) inhibitor. J. Med. Chem. 2011, 54, 6206–6214. [Google Scholar] [CrossRef]
  217. Badshah, H.; Ali, T.; Shafiq-ur, R.; Faiz-ul, A.; Ullah, F.; Kim, T.H.; Kim, M.O. Protective Effect of Lupeol Against Lipopolysaccharide-Induced Neuroinflammation via the p38/c-Jun N-Terminal Kinase Pathway in the Adult Mouse Brain. J. Neuroimmune Pharmacol. 2016, 11, 48–60. [Google Scholar] [CrossRef]
  218. Papi Reddy, K.; Singh, A.B.; Puri, A.; Srivastava, A.K.; Narender, T. Synthesis of novel triterpenoid (lupeol) derivatives and their in vivo antihyperglycemic and antidyslipidemic activity. Bioorganic Med. Chem. Lett. 2009, 19, 4463–4466. [Google Scholar] [CrossRef]
  219. Beattie, J.H.; Nicol, F.; Gordon, M.J.; Reid, M.D.; Cantlay, L.; Horgan, G.W.; Kwun, I.S.; Ahn, J.Y.; Ha, T.Y. Ginger phytochemicals mitigate the obesogenic effects of a high-fat diet in mice: A proteomic and biomarker network analysis. Mol. Nutr. Food Res. 2011, 55 (Suppl. S2), S203–S213. [Google Scholar] [CrossRef]
Figure 1. Landmark discoveries in the area of JNK research. The timeline of significant discoveries spans from the initial identification of JNKs as stress-activated protein kinases to the discoveries of selective JNK isoform-specific inhibitors.
Figure 1. Landmark discoveries in the area of JNK research. The timeline of significant discoveries spans from the initial identification of JNKs as stress-activated protein kinases to the discoveries of selective JNK isoform-specific inhibitors.
Biomolecules 14 00243 g001
Figure 2. Alternative splicing isoforms of the JNK MAPKs. (A) Characteristics of the alternative splicing isoforms of JNK1, JNK2, and JNK3. # indicates alternative exon 6 between the α and β isoforms of JNK1 (blue symbol “#”) or JNK2 (black symbol “#”) that result in a change of 5–7 amino acids in this region. Each of these JNK kinases is composed of a kinase domain (central region) as well as N- and C-termini represented by a line on the left (N terminus) and right (C terminus) of all kinase domains. Different colors and lengths in the N- or C-terminus represent distinct sequences and number of AAs compared to the main isoform. (B) The amino acid sequence of JNK1-α1, which contains 384 amino acids, is shown as a typical member of this family. JNK1 contains all 11 protein kinase subdomains as indicated (I–XI). The protein kinase activation loop is located between domains VII and VIII, which contains the conserved amino acid motifs DFG (red) and APE (red). The threonine (T, red) and tyrosine (Y, red) are within this loop and indicated by a red asterisk and a blue asterisk, respectively. (C) The structure of the human JNK1 protein (PDB id: 3ELJ). The N-/C-terminal lobes, N-terminal hairpin, and MAPK insert site as well as the activation loop are indicated. (D) Protein structural diagram of ten different isoforms of JNK produced by the alternative splicing of three JNK genes (JNK1, JNK2, and JNK3). The differences are indicated by the patterned rectangles. The similarity of the shorter forms is analyzed by comparing the percent identity of JNK1-β1, JNK2-α1, JNK2-β1, and JNK3-α1 with JNK1-α1. Similarly, the longer forms can be compared, and the percent identity of JNK1-α2 with JNK1-β2, JNK2-α2, JNK2-β2, and JNK3-α2 is shown. Red asterisk indicates the threonine, while blue asterisk indicates the tyrosine within the activation loop that is located between domains VII and VIII.
Figure 2. Alternative splicing isoforms of the JNK MAPKs. (A) Characteristics of the alternative splicing isoforms of JNK1, JNK2, and JNK3. # indicates alternative exon 6 between the α and β isoforms of JNK1 (blue symbol “#”) or JNK2 (black symbol “#”) that result in a change of 5–7 amino acids in this region. Each of these JNK kinases is composed of a kinase domain (central region) as well as N- and C-termini represented by a line on the left (N terminus) and right (C terminus) of all kinase domains. Different colors and lengths in the N- or C-terminus represent distinct sequences and number of AAs compared to the main isoform. (B) The amino acid sequence of JNK1-α1, which contains 384 amino acids, is shown as a typical member of this family. JNK1 contains all 11 protein kinase subdomains as indicated (I–XI). The protein kinase activation loop is located between domains VII and VIII, which contains the conserved amino acid motifs DFG (red) and APE (red). The threonine (T, red) and tyrosine (Y, red) are within this loop and indicated by a red asterisk and a blue asterisk, respectively. (C) The structure of the human JNK1 protein (PDB id: 3ELJ). The N-/C-terminal lobes, N-terminal hairpin, and MAPK insert site as well as the activation loop are indicated. (D) Protein structural diagram of ten different isoforms of JNK produced by the alternative splicing of three JNK genes (JNK1, JNK2, and JNK3). The differences are indicated by the patterned rectangles. The similarity of the shorter forms is analyzed by comparing the percent identity of JNK1-β1, JNK2-α1, JNK2-β1, and JNK3-α1 with JNK1-α1. Similarly, the longer forms can be compared, and the percent identity of JNK1-α2 with JNK1-β2, JNK2-α2, JNK2-β2, and JNK3-α2 is shown. Red asterisk indicates the threonine, while blue asterisk indicates the tyrosine within the activation loop that is located between domains VII and VIII.
Biomolecules 14 00243 g002
Figure 3. Crystal structures of JNKs and their ligands. (A) Crystal structure of JNK1 bound to a MKK7 docking motif (PDB ID: 4UX9). The D2 peptide (QRPRPTLQLPLA), which is derived from the D2 docking site of MKK7, is the bound ligand of JNK1. (B) Crystal structure of JNK1 in complex with the activating transcription factor 2 (ATF2) fragment (PDB ID: 6ZR5). Cyclic AMP-dependent transcription factor ATF2 (19-58) is the bound ligand of JNK1. (C) Crystal structure of AMP-bound human JNK2 (PDB ID: 7N8T). Adenosine monophosphate (AMP) is the bound ligand of JNK2. (D) Crystal structure of JNK3 complexed with a thiophenyl–pyrazolourea derivative (PDB ID: 7KSI). The thiophenyl–pyrazolourea derivative is the bound ligand of JNK3. All the four structural images were down-loaded from https://www.rcsb.or/. Access date: 28 October 2023.
Figure 3. Crystal structures of JNKs and their ligands. (A) Crystal structure of JNK1 bound to a MKK7 docking motif (PDB ID: 4UX9). The D2 peptide (QRPRPTLQLPLA), which is derived from the D2 docking site of MKK7, is the bound ligand of JNK1. (B) Crystal structure of JNK1 in complex with the activating transcription factor 2 (ATF2) fragment (PDB ID: 6ZR5). Cyclic AMP-dependent transcription factor ATF2 (19-58) is the bound ligand of JNK1. (C) Crystal structure of AMP-bound human JNK2 (PDB ID: 7N8T). Adenosine monophosphate (AMP) is the bound ligand of JNK2. (D) Crystal structure of JNK3 complexed with a thiophenyl–pyrazolourea derivative (PDB ID: 7KSI). The thiophenyl–pyrazolourea derivative is the bound ligand of JNK3. All the four structural images were down-loaded from https://www.rcsb.or/. Access date: 28 October 2023.
Biomolecules 14 00243 g003
Figure 4. The general activation mechanism of the JNK signaling pathway. JNK kinases are activated by a series of phosphorylation events in response to extracellular stimuli such as growth factors, cytokines, environmental stresses, pathogens, GPCR agonists, and drugs. These stimuli first result in activation of the STE20 protein homologue germinal center kinases, including HPK, NIK, and GCK, which interact with and phosphorylate MEKK1, or directly cause the activation of membrane proximal MAP3K kinases such as ASK1, TAK1, or MLK3, which then phosphorylates and activates MKK4 and MKK7. The activated MKK4 or MKK7 further phosphorylates and activates distinct JNK isoforms through dual phosphorylation on tyrosine and threonine residues within a conserved tripeptide motif. Notably, DUSPs and PP2A can dephosphorylate JNK and inhibit JNK signaling. Upon activation, JNK, on the one hand, phosphorylates several nuclear transcription factors such as c-Jun, ATF2, and SP1, causing the transcription of target genes. On the other hand, JNK also interacts with and phosphorylates a number of non-nuclear proteins, including mitochondrial BCL-2 family members (e.g., BCL-XL and BCL-2) and some cytoplasmic proteins. Finally, these proteins mediate a variety of cellular responses, including cell growth, differentiation, proliferation, survival, apoptosis, inflammation, metabolic stress, and cell death. In addition, several protein complexes are also involved in regulating JNK signaling. For instance, the complex composed of ACID, JIP1, and JNK1, the complex composed of JIP1 and JNK1/2/3, and the complex composed of Sab and JNK1/2/3 are also involved in JNK signaling transduction and are critical for the development of neurodegenerative diseases, autoimmune inflammatory diseases/glomerulosclerosis, and non-alcoholic steatohepatitis, respectively. The TAT-JIP1 peptide and Notch 1-IC are used for ameliorating autoimmune inflammatory diseases and glomerulosclerosis, respectively.
Figure 4. The general activation mechanism of the JNK signaling pathway. JNK kinases are activated by a series of phosphorylation events in response to extracellular stimuli such as growth factors, cytokines, environmental stresses, pathogens, GPCR agonists, and drugs. These stimuli first result in activation of the STE20 protein homologue germinal center kinases, including HPK, NIK, and GCK, which interact with and phosphorylate MEKK1, or directly cause the activation of membrane proximal MAP3K kinases such as ASK1, TAK1, or MLK3, which then phosphorylates and activates MKK4 and MKK7. The activated MKK4 or MKK7 further phosphorylates and activates distinct JNK isoforms through dual phosphorylation on tyrosine and threonine residues within a conserved tripeptide motif. Notably, DUSPs and PP2A can dephosphorylate JNK and inhibit JNK signaling. Upon activation, JNK, on the one hand, phosphorylates several nuclear transcription factors such as c-Jun, ATF2, and SP1, causing the transcription of target genes. On the other hand, JNK also interacts with and phosphorylates a number of non-nuclear proteins, including mitochondrial BCL-2 family members (e.g., BCL-XL and BCL-2) and some cytoplasmic proteins. Finally, these proteins mediate a variety of cellular responses, including cell growth, differentiation, proliferation, survival, apoptosis, inflammation, metabolic stress, and cell death. In addition, several protein complexes are also involved in regulating JNK signaling. For instance, the complex composed of ACID, JIP1, and JNK1, the complex composed of JIP1 and JNK1/2/3, and the complex composed of Sab and JNK1/2/3 are also involved in JNK signaling transduction and are critical for the development of neurodegenerative diseases, autoimmune inflammatory diseases/glomerulosclerosis, and non-alcoholic steatohepatitis, respectively. The TAT-JIP1 peptide and Notch 1-IC are used for ameliorating autoimmune inflammatory diseases and glomerulosclerosis, respectively.
Biomolecules 14 00243 g004
Figure 5. Molecular mechanisms for the involvement of the JNK pathway in insulin resistance. The JNK1/2 kinases were important for obesity-driven insulin resistance through four mechanisms: (A) In insulin-target cells, JNK1/2 directly phosphorylates IRS1 and IRS2 at serine and threonine residues, resulting in reduced tyrosine phosphorylation of IRS1/2 molecules, which further weakens the response of the PI3K/AKT signaling pathway to insulin. (B) In adipose tissue macrophages, JNK1/2 phosphorylates some transcriptional factors such as c-JUN, ATF2, and ELK1, stimulating gene transcription of pro-inflammatory cytokines, leading to increased levels of inflammatory “M1” cytokines (e.g., IL-6), which ultimately drive insulin resistance. (C) In pituitary thyrotropic cells, JNK1/2 phosphorylates c-JUN and c-FOS, sustaining the transcription and subsequent expression of DIO2, which mediates the shift from T4 to T3, finally leading to a T3-dependent reduction of TSHβ production. The reduced TSH ultimately causes low circulating levels of thyroid hormone, increases metabolic efficiency, and increases adiposity, finally driving insulin resistance. (D) In hepatocytes, JNK1/2 phosphorylates c-JUN, sustaining the transcription and expression of the transcription co-repressor NcoR1. Then, NcoR1 represses PPARα-driven gene expression, causing a reduction in FGF21 production, fatty acid oxidation, and ketogenesis, ultimately promoting fatty liver and insulin resistance.
Figure 5. Molecular mechanisms for the involvement of the JNK pathway in insulin resistance. The JNK1/2 kinases were important for obesity-driven insulin resistance through four mechanisms: (A) In insulin-target cells, JNK1/2 directly phosphorylates IRS1 and IRS2 at serine and threonine residues, resulting in reduced tyrosine phosphorylation of IRS1/2 molecules, which further weakens the response of the PI3K/AKT signaling pathway to insulin. (B) In adipose tissue macrophages, JNK1/2 phosphorylates some transcriptional factors such as c-JUN, ATF2, and ELK1, stimulating gene transcription of pro-inflammatory cytokines, leading to increased levels of inflammatory “M1” cytokines (e.g., IL-6), which ultimately drive insulin resistance. (C) In pituitary thyrotropic cells, JNK1/2 phosphorylates c-JUN and c-FOS, sustaining the transcription and subsequent expression of DIO2, which mediates the shift from T4 to T3, finally leading to a T3-dependent reduction of TSHβ production. The reduced TSH ultimately causes low circulating levels of thyroid hormone, increases metabolic efficiency, and increases adiposity, finally driving insulin resistance. (D) In hepatocytes, JNK1/2 phosphorylates c-JUN, sustaining the transcription and expression of the transcription co-repressor NcoR1. Then, NcoR1 represses PPARα-driven gene expression, causing a reduction in FGF21 production, fatty acid oxidation, and ketogenesis, ultimately promoting fatty liver and insulin resistance.
Biomolecules 14 00243 g005
Figure 6. Pro-oncogenic and anti-oncogenic roles of JNK signaling in cancers. (A) The JNK signaling plays a dual role in the occurrence and progression of different kinds of cancer, including hepatocellular carcinoma, multiple myeloma, prostate cancer, and oral cancer. (B) JNK signaling promotes the progression of glioblastoma and non-small-cell lung cancer. (C) Deficiency or inhibition of different JNKs promotes tumor formation in breast cancer, intestinal tumors, and papilloma. (D) Inhibition of the JNK pathway suppresses the tumorigenesis of ovarian cancer, skin cancer, and Burkitt lymphoma. Activation indicates sustained activation of the JNK pathway; inactivation represents deficiency of JNK1, JNK2, JNK3, or compound deficiency of JNK1 and JNK2; inactivation also indicates inhibition of JNK activity by JNK inhibitors such as SP600125.
Figure 6. Pro-oncogenic and anti-oncogenic roles of JNK signaling in cancers. (A) The JNK signaling plays a dual role in the occurrence and progression of different kinds of cancer, including hepatocellular carcinoma, multiple myeloma, prostate cancer, and oral cancer. (B) JNK signaling promotes the progression of glioblastoma and non-small-cell lung cancer. (C) Deficiency or inhibition of different JNKs promotes tumor formation in breast cancer, intestinal tumors, and papilloma. (D) Inhibition of the JNK pathway suppresses the tumorigenesis of ovarian cancer, skin cancer, and Burkitt lymphoma. Activation indicates sustained activation of the JNK pathway; inactivation represents deficiency of JNK1, JNK2, JNK3, or compound deficiency of JNK1 and JNK2; inactivation also indicates inhibition of JNK activity by JNK inhibitors such as SP600125.
Biomolecules 14 00243 g006
Table 1. Summary of JNK inhibitors and their application in human disease models.
Table 1. Summary of JNK inhibitors and their application in human disease models.
ClassDrugApplication in Disease ModelReference
ATP-competitive small-molecule inhibitorsSP600125Ischemia/reperfusion; cancer; T2D; acute kidney injury; inflammation; viral infections; and sepsis-induced lung injury[27,181,185]
CEP-1347parkinson’s disease and cancer[41,45]
AS601245
(bentamapimod)
Ischemia-caused neuronal damage; cancer; and rheumatoid arthritis (RA)[1,186]
SR-3576Neurodegenerative diseases[28]
CC-930Renal/liver fibrosis[41]
CC-401Renal/liver injury[41]
ATP-non-competitive small-molecule inhibitorsBI-78D3Liver damage and type 2 diabetes[29,199]
Compound 9Type 2 diabetes[189]
Small peptide inhibitorsD-JNKI-1
(XG-102 or AM-111)
Neuron damage; lipid metabolism; type 2 diabetes; cerebral ischemia; hearing loss; hypoxia-induced retinopathy; liver injury; and acute inflammatory insult[27,68,190,191,200]
IB1/2Type 2 diabetes[190,193]
Natural phytochemicalsC66Diabetes; aortic damage; and atherosclerosis[27,201,202,203]
LupeolNeuroinflammation; diabetes; and lipid metabolism[198,199]
GingerolLipid metabolism[200]
CapsaicinInflammatory insult[26]
Table 2. The specificity and structural information of several JNK inhibitors.
Table 2. The specificity and structural information of several JNK inhibitors.
InhibitorSpecificityOff-Target EffectsStructure of JNK Complexed with Inhibitor (PDB ID)
SP600125JNK1/2/3Yes1UKH
CEP-1347MLKYesNot reported
AS601245JNK1/2YesNot reported
SR-3537JNK3Yes3FI3
SR-3576JNK3Not reportedNot reported
CC-359JNK1/2/3Not reported3TTJ
CC-930JNK1/2/3Not reported3TTI
BI-78D3JNK/JIP1Not reportedNot reported
D-JNKI-1JNK1Not reportedNot reported
JIP1 peptideJNK1/3Not reported4H39; 1UKI
C66JNK2YesNot reported
LupeolJNK1YesNot reported
GingerolJNK1/2YesNot reported
CapsaicinJNK1YesNot reported
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

Yan, H.; He, L.; Lv, D.; Yang, J.; Yuan, Z. The Role of the Dysregulated JNK Signaling Pathway in the Pathogenesis of Human Diseases and Its Potential Therapeutic Strategies: A Comprehensive Review. Biomolecules 2024, 14, 243. https://doi.org/10.3390/biom14020243

AMA Style

Yan H, He L, Lv D, Yang J, Yuan Z. The Role of the Dysregulated JNK Signaling Pathway in the Pathogenesis of Human Diseases and Its Potential Therapeutic Strategies: A Comprehensive Review. Biomolecules. 2024; 14(2):243. https://doi.org/10.3390/biom14020243

Chicago/Turabian Style

Yan, Huaying, Lanfang He, De Lv, Jun Yang, and Zhu Yuan. 2024. "The Role of the Dysregulated JNK Signaling Pathway in the Pathogenesis of Human Diseases and Its Potential Therapeutic Strategies: A Comprehensive Review" Biomolecules 14, no. 2: 243. https://doi.org/10.3390/biom14020243

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop