Next Article in Journal
Langerhans Cells in Sentinel Lymph Nodes from Melanoma Patients
Previous Article in Journal
Effects of KEAP1 Silencing on NRF2 and NOTCH Pathways in SCLC Cell Lines
Previous Article in Special Issue
The Important Role of Protein Kinases in the p53 Sestrin Signaling Pathway
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

“Oh, Dear We Are in Tribble”: An Overview of the Oncogenic Functions of Tribbles 1

by
Karnika Singh
,
Christian A. Showalter
,
Heather R. Manring
,
Saikh Jaharul Haque
and
Arnab Chakravarti
*
Department of Radiation Oncology, The Ohio State University Comprehensive Cancer Center, Columbus, OH 43210, USA
*
Author to whom correspondence should be addressed.
Cancers 2024, 16(10), 1889; https://doi.org/10.3390/cancers16101889
Submission received: 18 April 2024 / Revised: 11 May 2024 / Accepted: 13 May 2024 / Published: 16 May 2024
(This article belongs to the Special Issue Protein Kinases and Pseudokinases in Cancers)

Abstract

:

Simple Summary

Pseudokinases have evolved from conventional kinases and differ from them in their inability to phosphorylate their substrates. The Tribbles pseudokinases have evolved from the CAMK family of kinases and include three members: Tribbles 1, 2, and 3. Tribbles 1, which evolved later in metazoan lineage, appears to have a regulatory function. It interacts with a variety of substrates through different domains embedded in its molecular structure. Each interaction regulates cellular processes involved in cell division, survival, and metabolism. These processes also extrapolate to cancer development and other diseases. Based on its involvement in disorders and therapy resistance, it can be considered a potential candidate for drug development. The improved knowledge about its function can be utilized to design small-molecule inhibitors against TRIB1.

Abstract

Pseudokinases are catalytically inactive proteins in the human genome that lack the ability to transfer phosphate from ATP to their substrates. The Tribbles family of pseudokinases contains three members: Tribbles 1, 2, and 3. Tribbles 1 has recently gained importance because of its involvement in various diseases, including cancer. It acts as a scaffolding protein that brings about the degradation of its substrate proteins, such as C/EBPα/β, MLXIPL, and RAR/RXRα, among others, via the ubiquitin proteasome system. It also serves as an adapter protein, which sequesters different protein molecules and activates their downstream signaling, leading to processes, such as cell survival, cell proliferation, and lipid metabolism. It has been implicated in cancers such as AML, prostate cancer, breast cancer, CRC, HCC, and glioma, where it activates oncogenic signaling pathways such as PI3K-AKT and MAPK and inhibits the anti-tumor function of p53. TRIB1 also causes treatment resistance in cancers such as NSCLC, breast cancer, glioma, and promyelocytic leukemia. All these effects make TRIB1 a potential drug target. However, the lack of a catalytic domain renders TRIB1 “undruggable”, but knowledge about its structure, conformational changes during substrate binding, and substrate binding sites provides an opportunity to design small-molecule inhibitors against specific TRIB1 interactions.

1. Introduction

The human genome encodes a total of 581 protein kinases, of which 50 are catalytically inactive or weak and termed pseudokinases because they lack at least one residue in their conserved catalytic site [1]. Some pseudokinases retaining weak kinase activity include ErbB3/HER3, in which a couple of amino acids in its catalytic domain are replaced by other amino acids (Glu740> His and Asp815> Asn), but it is still able to bind ATP and catalyze phosphoryl transfer with lower efficiency [2]. Pseudokinases are known to perform a variety of physiological functions, such as allosteric regulation of other enzymes (e.g., VRK3 [3]), scaffolding of signaling components (e.g., Tribbles family [4]), and acting as molecular switches (e.g., MLKL [5]). Many pseudokinases are implicated in several diseases, such as cancer, metabolic disorders, and autoimmune diseases, among others [6]. One such pseudokinase is mammalian tribbles homolog 1 (TRIB1), a member of the Tribbles family of pseudokinases, which has emerged to play a role in several diseases, including cancer, and is the focus of this review.

2. Structure and Function of TRIB1

TRIB1 is encoded by the TRIB1 gene on chromosome 8 (8q24.13 locus) in the human genome. It is also known as C8FW, TRB1, GIG2, and SKIP1. TRIB1 was first identified in 1997 by Wilkin and colleagues (human clone C8FW), and in spite of a ~30% sequence identity with the protein kinase family, they did not observe an ATP binding site in their isolated clone [7]. The half-life of TRIB1 mRNA is less than 1 h [8], and protein turnaround is about 2 h [9]. The TRIB1 protein has two isoforms, isoforms 1 and 2. The isoform 1 encoded by the longer transcript is the full-length protein (41 kDa), whereas the translation of isoform 2 is initiated at a downstream start codon, and the encoded protein contains a shorter N-terminus (23.48 kDa) [10]. TRIB1 is highly expressed in the bone marrow, liver, thyroid gland, urinary bladder, adipose tissue, gall bladder, and parathyroid gland [11]. TRIB1 protein is found in both the cytoplasmic [12] and nuclear [13] compartments of the cell. Several transcription factor binding sites are predicted to be present in the TRIB1 gene promoter 5′ UTR region, including E2F, C/EBPβ, and NF-1 [4]. Similarly, many miRNAs also putatively bind to the 3′UTR of TRIB1 mRNA and regulate its gene expression and function [14]. ATF3 regulates TRIB1 gene expression in response to proteasome inhibition in hepatocyte cell lines [15].

2.1. Evolution and Structure of TRIB1

TRIB1 belongs to the Tribbles family of pseudokinases, comprised of TRIB1, TRIB2, TRIB3, and a distantly related member, STK40. The name Tribbles comes from their homology with the domain structure of the Drosophila tribbles protein [16]. In 2000, TRIB1 was first described to have a function during Drosophila gastrulation in the modulation of String/Cdc25 proteins in mesoderm cells [16,17,18]. TRIB1 orthologs are mainly encountered in later metazoan lineages (the vertebrates), as opposed to TRIB2, which is also observed in the oldest metazoans [4]. This is suggestive of the fact that TRIB1 evolved later in evolution, possibly to accommodate complex cell signaling processes in higher organisms.
The structure of TRIB1 is composed of three domains: The N-terminal domain (AA 1–90), the central pseudokinase domain (AA 91–330), and the C-terminal domain (AA 331–373) [19,20]. The N-terminal domain of TRIB1 contains a putative nuclear localization signal spanning residues 33–51, which is shown to be required for nuclear localization [9]. The N-terminal domain is also rich in proline (P), glutamate (E), serine (S), and threonine (T) residues (AA 53–88), which are characteristic of PEST domains; however, it is also shown to be unassociated with TRIB1 instability [15]. The pseudokinase domains of TRIB1, 2, and 3 share sequence similarity with the canonical kinase domain of CAMK (calcium/calmodulin-dependent protein kinase) [21]. Furthermore, at the amino acid level, TRIB1 shares 71% sequence similarity with TRIB2 and 54% sequence similarity with TRIB3 in its pseudokinase domain region. The structural features of the pseudokinase domain that distinguish it from a canonical kinase include a bent αC helix and divergent glycine-rich loop, both of which are required for ATP binding in a bona fide protein kinase. The Asp-Phe-Gly (DFG) motif is replaced by the Ser-Leu-Glu (SLE) sequence that occludes the nucleotide binding pocket. All of these structural/sequence changes represent a deformed N-terminal lobe and prevent the binding of ATP to TRIB1, thus lacking or inhibiting any phosphoryl transfer, which in turn classifies it as a pseudokinase [22]. The pseudokinase domain also allows for binding of interacting partners and substrates, with C/EBPα/β (AA 168–293 [23]) being its most characterized substrate so far. The C-terminal domain of TRIB1 houses two important protein–protein interaction motifs: ILLHPWF (AA 332–339 [12]) that binds MEK1/2, and DQIVPE (AA 355–360 [24]) for binding COP1 E3 ubiquitin ligase. It has been shown by Murphy et al. that there is an intramolecular interaction between the C-terminal tail of TRIB1 and its pseudokinase domain, which in turn increases the stability of the TRIB1 protein [22]. This interaction is also observed in the CAMK group of protein kinases, where the C-terminal regulatory region makes multiple inhibitory interactions with its catalytic core to play an autoinhibitory role [25]. Both TRIB2 and TRIB3 also contain a MEK1/2 binding site, I(L/D)LHPW(F/L) (AA 30(2/9)-3(07/15)), and a COP1 binding site (D/A)Q(L/V)VPD (AA 3(26/33)-3(31/37)) on their C-termini [26]. Overall, TRIB2 (343 amino acids) and TRIB3 (358 amino acids) are smaller proteins than TRIB1, with a shorter N-terminal domain as revealed by amino acid alignment.

2.2. Functions of TRIB1

TRIB1 has been recognized as an adapter protein that provides a scaffold for the degradation of its substrates (summarized in Table 1). The most characterized substrate of TRIB1 is C/EBPα, a lineage-specific transcription factor that undergoes degradation via COP1 E3 ubiquitin ligase [22]. Binding of C/EBPα to the TRIB1 pseudokinase domain initiates a conformational change in the SLE motif to release the COP1 binding motif on the C-terminal domain, in turn relieving the autoinhibition. This substrate-bound conformation of TRIB1 is known as the “SLE-in” conformation and corresponds to the active protein kinase conformation [23]. COP1 then binds to the C-terminus through its WD40 domain [24] and causes ubiquitination of C/EBPα at its lysine residues, ultimately causing its degradation through the ubiquitin proteasome system [27]. After the degradation of C/EBPα, the C-terminal domain folds itself back onto the pseudokinase domain, and the autoinhibitory conformation is restored. This conformation is known as “SLE-out” and is similar to an inactive protein kinase [28]. The tribbles protein also controls the protein levels of slbo, the Drosophila homolog of the C/EBP family of transcription factors, via degradation and inhibits border cell migration during oogenesis [29]. Furthermore, because the C/EBP family of transcription factors determines the switch between granulopoiesis and monopoiesis [30], degradation of C/EBPα by TRIB1 regulates myeloid cell differentiation [31]. TRIB1 has also been shown to negatively regulate retinoic acid receptor (RAR) signaling by forming a complex with it and its binding partner RXR (retinoid X receptor) [32].
The other primary function of TRIB1 is the sequestration of signaling molecules and subsequent activation of the downstream pathways (summarized in Table 1). As mentioned above, the pseudokinase and the C-terminal domains of TRIB1 contain sites that bind to Akt and MEK1/2, respectively. Drosophila tribbles have been shown to interact with Akt1 through its kinase-like domain and inhibit its activation [33]. Another report shows that TRIB1 binds to Akt, possibly at its pseudokinase domain, and causes its activation [34]. TRIB3 has also been shown to interact with Akt1 through its pseudokinase domain [35]. The mediation of Akt activation and its downstream signaling by TRIB1 suggests that TRIB1 is a modulator of the PI3K-Akt pathway [36]. Alternatively, TRIB1 also interacts with MEK1/2 through its C-terminal domain and regulates the activation of its downstream protein, ERK [37]. TRIB1 also interacts with MKK4 through its pseudokinase domain and modulates vascular smooth muscle cell proliferation and chemotaxis by activating the downstream JNK pathway [38]. TRIB1 also plays a role during M1 and M2 macrophage polarization of murine bone marrow-derived macrophages (BMDMs) by modulating the JAK/STAT pathway [39]. TRIB1 has been shown to be important for the differentiation of tissue-resident M2-like macrophages (F4/80+ MR+), which are involved in the maintenance of adipose tissues [31].
Table 1. The interacting partners of TRIB1.
Table 1. The interacting partners of TRIB1.
Interacting PartnerInteraction SiteFunction/Effect References
C/EBPα/β Pseudokinase domain (AA168–293)Degradation of C/EBPα/β [22,23]
COP1C-terminal domain (AA 355–360)Proteasome-mediated degradation of substrates such as C/EBPα/β[22,24]
MEK1/2C-terminal domain
(AA 332–339)
MAPK pathway activation[12,37]
RAR/RXRαPseudokinase domainDownregulation of RAR signaling[32]
AktPseudokinase domain (AA 90–160)Cell survival through the activation of PI3K-Akt pathway[34,36]
MKK4Pseudokinase domainJNK pathway activation, vascular smooth muscle cell proliferation and chemotaxis[38]
MLXIPL (ChREBP)Pseudokinase domainDegradation of MLXIPL (ChREBP) leading to transcriptional inhibition of lipogenesis genes[40]
HNF4APseudokinase domain, multiple epitopes but AA 1–240 exhibit strongest bindingGene expression of lipid metabolism and lipoprotein regulators and genes in liver and intestine[41]
SAP18/Sin3ACOP1 binding siteModulation of MTTP expression involved in lipid metabolism[42]
Nrf2C-terminal domainNrf2 sequestration in cytoplasm suppressing liver regeneration[43]
NF-кBN-terminal and Pseudokinase domainDirect promoter recruitment to NF-кB DNA recognition site and upregulation of cytokine gene expression[44]
FoxP3N-terminal and Pseudokinase domainTregs regulation[45]
FERMT2Pseudokinase domainDegradation of FERMT2 and transcriptional regulation through β-catenin[46]
ZBT7AUnknownModulation of ER-associated transcription[46]
CUL4A/BC-terminal domainDegradation of substrates[46]
HDAC1UnknownP53 deacetylation[47]
CD72Pseudokinase domainDegradation of CD72 leading to development of autoimmunity[48]
MALT1N-terminal domain (AA 83–89)T cell receptor signaling regulation[49]

2.2.1. TRIB1 and Lipid Metabolism

TRIB1 is known to play an important role in lipid metabolism. It was first identified to be associated with triglycerides, LDL, and HDL cholesterol in a genome-wide association study (GWAS) [50]. TRIB1 knockout mice exhibit increased triglyceride and plasma cholesterol levels, which are accompanied by downregulation of lipogenic genes such as acetyl-CoA carboxylase and fatty acid synthase, among others [51]. Liver-specific deletion of TRIB1 has revealed that elevated levels of C/EBPα increase the transcription of fatty acid synthesis genes and result in pathologic phenotypes, suggesting a role of TRIB1 in keeping the transcription of lipogenic genes in check via degradation of C/EBPα [52,53]. However, on the contrary, a recent report suggests that adipocyte specific knockdown of TRIB1 results in reduced plasma triglyceride and cholesterol levels along with increased adiponectin secretion, which are associated with improved metabolic health in humans [54]. TRIB1 also interacts with the hepatic lipogenic master regulator MLXIPL (MLX interacting protein like), also known as ChREBP, and causes its degradation, leading to transcriptional inhibition of genes involved in liponeogenesis [40]. The physical interaction of TRIB1 with HNF4A has also been reported, which plays a role in liver and intestinal gene expression as well as lipid metabolism [41]. TRIB1 also acts as a scaffold for the SAP18 (Sin3A-associated protein) and mSin3A interactions, which modulate the expression of MTTP (microsomal TG transfer protein) in the mouse liver, ultimately impacting lipid metabolism [42]. TRIB1 also has the capability to suppress liver regeneration by modulating redox homeostasis in hepatocytes via blockade of Nrf2 nuclear translocation [43]. TRIB1 genetic loci are also associated with liver function in terms of certain enzymes’ secretion, such as alanine transaminase (ALT), alkaline phosphatase (ALP), and γ-glutamyl transferase (GGT), which are commonly used as markers of certain liver ailments, suggesting it to be a genetic trait that might be inherited among humans [55].

2.2.2. TRIB1 and Innate Immunity

TRIB1 regulates NF-кB through various mechanisms, namely, regulation of p100 expression and inhibition of IKKα via phosphorylation by Akt on T23, among others. Regulation of NF-кB by TRIB1 is supported by TRIB1 knockdown studies, which resulted in downregulation of NF-кB-transcribed genes such as CCND1 (cyclin D1) and IL-8 (interleukin 8), suggesting a role of TRIB1 in cell cycle transition and inflammation [36]. Furthermore, the IL-8 promoter contains AP-1 and C/EBPα binding sites, both of which have been shown to be regulated by TRIB1 [19]. TRIB1 is also shown to interact with the RelA subunit of NF-кB in white adipocytes and acts as its transcriptional co-activator to regulate client cytokine gene expression [44]. Interaction of TRIB1 with FoxP3, the master regulator of regulatory T cells (Tregs), is reported in human adherent cell lines and Tregs [45,56]. TRIB1 also increases IL-2 production in activated T cells through NFAT (nuclear factor of activated T cells) [57]. TRIB1-mediated downregulation of C/EBPβ results in altered TLR-mediated signaling in macrophages, especially in the context of C/EBPβ-induced gene expression [58]. Knockdown of TRIB1 in macrophages inhibits their migration and increases production of TNFα [59]. TRIB1 is identified as a potential biomarker for chronic antibody-mediated allograft rejection due to its increased mRNA levels present in peripheral blood mononuclear cells of patients with deteriorating renal graft function [60]. Taken together, these reports suggest that TRIB1 may have a function in innate immunity as well. The functional overview of TRIB1 is outlined in Figure 1.

3. Role of TRIB1 in Disease

Based on its various functions described above, TRIB1 has been implicated in numerous diseases and malignancies. Its close proximity to MYC and its involvement in fundamental pathways such as MAPK signaling, NF-кB signaling, and the PI3K-Akt pathway have led to the identification of TRIB1 as an oncogene that plays a pivotal role in tumor progression and maintenance. Furthermore, its involvement in lipid metabolism and innate immunity causes other disease modalities such as atherosclerosis and inflammation.

3.1. TRIB1 and Cancer

3.1.1. Acute Myeloid Leukemia (AML)

It was observed in an AML patient that TRIB1 was overexpressed along with MYC, both of which reside on the same chromosomal region (8q24) 2.25 Mb apart [61]. AML is characterized by the accumulation of hematopoietic progenitor cells in the bloodstream due to genetic alterations affecting their differentiation [62]. As mentioned earlier, C/EBPα is the lineage-specific transcription factor that is targeted for degradation by TRIB1 [22,30]. It was later discovered that overexpressed TRIB1 drives Hoxa9-induced leukemogenesis by decreasing C/EBPα protein levels, thereby modulating the enhancer programs at the Erg and Spns2 loci [63]. Hoxa9 and Meis1 are normally expressed in early progenitor cells and decrease in expression upon their terminal differentiation [64]. In the context of AML, Hoxa9 and Meis1 are overexpressed in primary bone marrow cells and cause leukemogenesis [65,66,67]. Hoxa9 is a regulator of the primitive hemopoietic compartment, and Meis1 is a transcription factor critical for leukemia stem cell regulation [68]. Specific interaction between TRIB1 and the coactivation of Hoxa9/Meis1 is reported in AML [65]. TRIB1-mediated activation of the MAPK pathway also plays a role in AML by enhancing the self-renewal of malignant bone marrow cells [12].

3.1.2. Prostate Cancer

TRIB1 and cMYC are observed to be co-amplified in prostate cancer as well. TRIB1 mRNA levels are also similarly elevated in the prostate of PTEN null mice, exhibiting increased proliferation potential [69]. TRIB1 overexpression causes increased sphere formation in prostate cancer cell lines and increased tumor formation in a xenograft mouse model [70]. TRIB1 influences the prostate cancer tumor microenvironment by causing CD136+ macrophage infiltration, promoting M2 macrophage differentiation, and affecting cytokine secretion by inhibiting IкB-ζ (NF-кB inhibitor zeta) [71].

3.1.3. Breast Cancer

TRIB1 and cMYC co-amplification is further observed in breast cancer patients [46] and associated with reduced overall and disease-free survival [46]. TRIB1 regulates both the G1/S [36] and G2/M transition [46] in breast cancer cells. A recent study utilizing the qPLEX-RIME technique [72] has revealed that TRIB1 interacts with the β-catenin co-factor FERMT2 (FERM Domain Containing Kindlin 2) and ER co-factor ZBTB7A (Zinc Finger and BTB Domain Containing 7A) in breast cancer cells, giving further insights into the TRIB1 interactome and its involvement in transcriptional regulation through β-catenin and ER [46]. TRIB1 reduces DR5 protein levels in breast cancer cells through its elevated NF-кB signaling, thus decreasing TRAIL-induced apoptosis [36]. Heightened expression of TRIB1 in tumor associated macrophages (TAMs) influences the breast cancer tumor microenvironment by regulating oncogenic cytokine expression [73].

3.1.4. Colorectal Cancer (CRC)

TRIB1 has been shown to be amplified in both CRC cell lines and patients [74]. TRIB1 and MYC are co-amplified in CRC cells. TRIB1 protein expression significantly correlates with the expression of proteins such as p-ERK, p-MEK, Akt, PTEN, cleaved caspase 3, and MYC in CRC cells [75]. TRIB1 overexpression correlates with poor overall survival in CRC patients. TRIB1 promotes migration and invasion of CRC cells through the activation of FAK/Src and ERK pathways, resulting in an upregulation of MMP-2 expression [76].

3.1.5. Other Cancers

The TRIB1 gene is also associated with pancreatic cancer in a Chinese Han population [77]. TRIB1 is observed to be highly expressed in bone marrow mononuclear cells of multiple myeloma (MM) patients with progressive disease compared to those in remission. Such patients have an increased percentage of M2 macrophages, suggesting that TRIB1 plays a role in M2 macrophage polarization in MM [78]. TRIB1 plays a role in hepatocellular carcinoma (HCC) as well, as evidenced by its upregulation in HCC cell lines and tumor tissues. TRIB1 promotes cell proliferation, migration, invasion, and epithelial-to-mesenchymal transition (EMT) in HCC cell lines. TRIB1 upregulation is accompanied by p53 downregulation and increased β-catenin signaling [79], as observed in breast cancer cells. A previous report suggests that TRIB1 can downregulate p53’s transcriptional activity by forming a complex with HDAC1, leading to HDAC1 catalyzed p53 deacetylation, thereby reducing its DNA binding activity [47]. Moreover, it has been shown in glioma and primary GBM cells that TRIB1 reduces p53 levels through a similar mechanism [34,80]. Increased TRIB1 mRNA levels have been correlated with poor overall survival of glioma patients [34]. This mechanism suggests that TRIB1 could play a central role in tumor initiation and maintenance by modulating the tumor suppressor activities of p53.

3.2. TRIB1 and Other Diseases

In addition to cancer, TRIB1 has been implicated in metabolic and lipid disorders, according to several GWAS. TRIB1 variants are directly associated with a reduced risk of myocardial infarction in humans [51]. Similarly, a previous review has discussed that TRIB1 is linked to the etiology of Crohn’s disease, non-alcoholic fatty liver disease, dyslipidemia, and cardiovascular disease [81]. Increased TRIB1 gene expression is encountered in the coronary arteries of patients with ischemic heart disease [38]. Genetic deletion of TRIB1 promotes atherosclerosis in mice, which is accompanied by pronounced hyperlipidemia and hepatic and systemic inflammation [82]. However, lack of TRIB1 in myeloid cells decreases early atheroma formation and reduces atherosclerosis burden in mouse models [83]. A recent meta-analysis reveals that the TRIB1 SNP (Rs2954029) A allele present in the 8q24 locus is positively associated with the risk of coronary artery disease (CAD) combined with stroke [84]. Several studies have also explored the association between TRIB1 and non-alcoholic liver disease and hepatic steatosis in terms of SNPs (rs6982502, rs2954021, rs17321515, rs2954029) in different populations [40,85,86,87]. Transcriptomics studies in similar mouse models reveal genes enriched in VLDL assembly, TG biosynthesis, gluconeogenesis, and glycogen synthesis pathways are upregulated after TRIB1 overexpression [40,85,86,87]. In order to exploit TRIB1’s ability to modulate lipid metabolism, Nagiec et al. identified a small molecule BRD0418, a benzofuran compound that increases TRIB1 expression in HepG2 cells, leading to reduced VLDL production, cholesterol biosynthesis, and increased LDL uptake [88]. The cytokine oncostatin M, which is shown to upregulate LDLR gene transcription in HepG2 cells [89], also acts by increasing TRIB1 gene expression [88]. Another tricyclic glycal compound, BRD8518, was later identified to have higher potency than BRD0418 in inducing TRIB1 and LDLR gene expression. This compound affects the MAPK pathway but is not suitable for in vivo testing because of its ADME profile [90]. TRIB1 is also suggested to influence obesity by regulating brown adipose tissue mitochondrial function via respiratory chain complex III [91]. Of note, TRIB1 is shown to negatively regulate B cells in systemic lupus erythematosus via its interaction with CD72, an effector of autoimmunity [48].

4. Role of TRIB1 in Cancer Therapy Resistance

Therapy resistance has always been considered the biggest roadblock in cancer treatment and relapse. Based on the above discussion, it is clear that TRIB1 plays an important role in the malignant transformation of various tumors and modulates several mechanisms for tumor sustenance. Several reports have shown that its involvement in various pathways, such as PI3K-AKT, MAPK, p53-HDAC1, and NF-кB, among others, suggests TRIB1 plays a role in cancer therapy resistance. TRIB1 was identified as an exclusive downstream effector of mutant PIK3CA in genetically modified lung epithelial cells [92]. A later report showed that downregulation of p53 by the TRIB1-HDAC1 interaction induces resistance to cisplatin chemotherapy and cancer stem cell enrichment in non-small cell lung carcinoma (NSCLC) [93]. Similarly, the TRIB1 gene was identified as part of a gene signature induced in breast cancer cell lines in response to long-term paclitaxel treatment, causing treatment resistance [94]. TRIB1 overexpression causes resistance to ATRA (all-trans retinoic acid) treatment during acute promyelocytic leukemia in exclusively sensitive myeloid cells expressing the PML/RARA fusion protein by preventing their differentiation through the downregulation of C/EBPα [95]. We and others have shown that TRIB1 mRNA and protein levels are upregulated by radiation and temozolomide treatment in glioma cells, causing a decrease in treatment-induced cell death [34,80]. In primary GBM cells, TRIB1 causes upregulation of the ERK and Akt pathways and modulation of p53 function through COP1 and HDAC1, further contributing to therapy resistance [34]. TRIB1 negatively regulates the anti-tumor cytokine IL-15 in TAMs in breast tumors, causing a decrease in T-cells that promote anti-tumor responses [73]. TRIB1 has been identified as a negative regulator of T-cell receptor (TCR) signaling via its interaction with MALT1, a TCR activator, leading to disruption of MALT1 signaling complexes responsible for optimal T cell activation and function [49]. Deletion of TRIB1 from T cells has been reported to improve PD-L1 immune checkpoint blockade, suggesting a role for TRIB1 in immunotherapy resistance [96]. A recent report has similarly shown that in a CRC mouse model, T cell recruitment to the tumor was improved after TRIB3 loss, thereby sensitizing the tumor to immune checkpoint blockade therapy [97]. The roles of TRIB1 in cancer and therapy resistance are further discussed in Table 2.

5. Potential Strategies to Target TRIB1

In light of the above discussion, it is clear that TRIB1 is involved in the tumor maintenance and therapy resistance of various cancers. Therefore, inhibiting TRIB1 can be considered a potential strategy for targeted cancer therapy that would alter the function/activity of important proteins such as Akt and MAPK in normal cells. In general, TRIB1 and many other pseudokinases have been considered “undruggable” for a very long time because, unlike conventional kinases, they do not possess an ATP binding domain, which has so far been the most druggable site in these proteins. On the contrary, the pseudokinases that possess weak catalytic activity (such as TRIB2, the JAK family, and HER3) have been targeted by small molecules against their catalytic site and demonstrated inhibition of their pseudokinase activities. There are a variety of FDA-approved JAK inhibitors for different diseases that act not only by targeting the ATP binding domain but also through allosteric inhibition [103]. Ruxolitinib was the first FDA-approved small-molecule inhibitor of JAK1/2 for the treatment of BCR-ABL1-negative myeloproliferative neoplasms and acts through competitive inhibition of the ATP-binding catalytic site on the kinase domain [104].
Given the variety of functions performed by TRIB1, several approaches can be utilized to disrupt its efficient operation. The most prominent function of TRIB1 is to act as a scaffold for the sequestration of signaling molecules, for which it adapts “SLE-in” (on) and “SLE-out” (off) conformations that allow for the substrates/effectors to bind. Small-molecule compounds locking TRIB1 in the SLE-in conformation would allow for blocking any substrate from binding, and locking in the SLE-out conformation would maintain TRIB1 in its auto-inhibitory conformation. Jamieson et al. have tested several compounds for their potential to stabilize TRIB1 in either of these conformations and reported moderate success [23]. Foulkes et al. reported that pre-existing EGFR inhibitors can target the low-affinity ATP binding site on TRIB2 and cause either its stabilization or destabilization [105]. Mutant K-Ras has been an “undruggable” target for years in lung cancer treatment, but the design of a covalent inhibitor that locks this protein in its “inactive conformation” has provided great promise in targeting this protein [106]. The recent resolution of the TRIB1 crystal structure and deeper understanding of the mechanisms of TRIB1 action have revealed new substrates and sites that can be used to design small-molecule inhibitors that are not only effective but are also blood–brain-barrier permeable.
Another potential approach for targeting TRIB1 could be the induction of its degradation through PROTACs (proteolysis targeting chimeras) and HyT (hydrophobic tagging) to reduce its levels inside the cell. PROTACs are bifunctional molecules that cause the degradation of a protein(s) of interest by bringing them within close proximity of an E3 ligase [107]. HyT, on the other hand, tags protein(s) of interest with hydrophobic fragments so that they are recognized as misfolded and damaged proteins and are therefore cleared through protein quality control mechanisms [108]. The pseudokinase Her3 has been targeted using both approaches. PROTAC selectively degrades HER3 but does not abrogate its downstream signaling [109], whereas HyT induces partial HER3 degradation and blocks HER3-dependent signaling in cell line models [110]. PROTAC 23 developed against IRAK3 pseudokinase exhibits degradation of around 98% IRAK3 protein in THP1 cells [111]. MLKL is another pseudokinase that has been successfully targeted using PROTAC 36 in the TSZ model of necroptosis, demonstrating an inhibition of cell death [112]. Other alternative approaches worth exploration include allosteric inhibition, antibodies, proteasome inhibitors, and irreversible inhibitors, among others.

6. Conclusions

TRIB1 plays a variety of roles in both health and disease. Although it may not phosphorylate any downstream substrates like a canonical kinase, it certainly acts as an intermediate for the sequestration of proteins to carry out important physiological functions such as differentiation, lipid metabolism, and innate immune function, among others. As was identified by Sanchez-Vega et al., there are at least ten cell signaling pathways that are found to be altered in different cancers [113]. Of these pathways, TRIB1 has been shown to play a role in the RTK/RAS, PI3K, Myc, p53, and cell cycle pathways in different tumor types, suggesting the importance of this pseudokinase in oncogenic signaling. TRIB1 and pseudokinases in general have only recently gained limelight in the cancer field for providing a new insight into tumorigenic processes. The implication of TRIB1 in malignant transformation also provides an opportunity for drug development for effective therapies. However, the lack of a catalytic domain on this protein may pose some hurdles, but its alternating “active” and “inactive” conformations hold potential as an attractive vulnerability to design effective drugs that would block its scaffolding function and disrupt downstream oncogenic signaling. PROTACs are an up-and-coming approach that have shown success in hitting undruggable targets and therefore warrants further exploration with regards to TRIB1. In conclusion, targeting TRIB1 presents an opportunity for cancer treatment as well as other metabolic disorders in which TRIB1 has been shown to play a role.

Author Contributions

Conceptualization, K.S. and S.J.H.; writing—original draft preparation, K.S. and C.A.S.; writing—review and editing, S.J.H. and H.R.M.; supervision, S.J.H. and H.R.M.; funding acquisition, A.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by: National Cancer Institute [R01CA169368 (to A.C.), R01CA11522358 (to A.C.), R01CA1145128 (to A.C.), R01CA108633 (to A.C.), R01CA188228 (to A.C., R.B., K.L. and J.B.), 1RC2CA148190 (to A.C.), and U10CA180850–01 (to A.C.)]; A Brain Tumor Funders Collaborative Grant (to A.C.), and Ohio State University Comprehensive Cancer Center Award (to A.C.).

Acknowledgments

The authors want to acknowledge Chunhua Han for her suggestions and technical assistance throughout this project. Her contribution to this study is deeply appreciated.

Conflicts of Interest

The authors declare no conflicts of interest.

List of Abbreviations

AAAmino Acid
AMLAcute Myeloid Leukemia
C/EBPαCAAT Enhancer Binding Protein α
CAMKCa2+/Calmodulin-Dependent Protein Kinase
ChREBPCarbohydrate Responsive Element Binding Protein
COP1Constitutive Photomorphogenic 1
EMTEpithelial to Mesenchymal Transition
EREstrogen Receptor
EGFREpidermal Growth Factor Receptor
Foxp3Forkhead box p3
HNF4AHepatocyte Nuclear Factor 4-Alpha
IRAK3Interleukin-1 receptor-associated kinase 3
JNKc-Jun N-terminal Kinase
MLKLMixed Lineage Kinase-Like
MLXMax like protein x
NF-кBNuclear factor kappa B
Nrf2Nuclear Factor Erythroid 2-Related Factor 2
MALT1Mucosa-Associated Lymphoid Tissue protein 1
PML/RARAPromyelocytic Leukemia/Retinoic Acid Receptor Alpha
TLRToll-like Receptor
HER3Human Epidermal Growth Factor Receptor 3
TRAILTNF-related apoptosis-inducing ligand
VRK3Vaccinia Related Kinase

References

  1. Manning, G.; Whyte, D.B.; Martinez, R.; Hunter, T.; Sudarsanam, S. The protein kinase complement of the human genome. Science 2002, 298, 1912–1934. [Google Scholar] [CrossRef]
  2. Shi, F.; Telesco, S.E.; Liu, Y.; Radhakrishnan, R.; Lemmon, M.A. ErbB3/HER3 intracellular domain is competent to bind ATP and catalyze autophosphorylation. Proc. Natl. Acad. Sci. USA 2010, 107, 7692–7697. [Google Scholar] [CrossRef]
  3. Kang, T.H.; Kim, K.T. Negative regulation of ERK activity by VRK3-mediated activation of VHR phosphatase. Nat. Cell Biol. 2006, 8, 863–869. [Google Scholar] [CrossRef]
  4. Eyers, P.A.; Keeshan, K.; Kannan, N. Tribbles in the 21st Century: The Evolving Roles of Tribbles Pseudokinases in Biology and Disease. Trends Cell Biol. 2017, 27, 284–298. [Google Scholar] [CrossRef]
  5. Sun, L.; Wang, H.; Wang, Z.; He, S.; Chen, S.; Liao, D.; Wang, L.; Yan, J.; Liu, W.; Lei, X.; et al. Mixed lineage kinase domain-like protein mediates necrosis signaling downstream of RIP3 kinase. Cell 2012, 148, 213–227. [Google Scholar] [CrossRef] [PubMed]
  6. Kung, J.E.; Jura, N. Prospects for pharmacological targeting of pseudokinases. Nat. Rev. Drug Discov. 2019, 18, 501–526. [Google Scholar] [CrossRef]
  7. Wilkin, F.; Savonet, V.; Radulescu, A.; Petermans, J.; Dumont, J.E.; Maenhaut, C. Identification and characterization of novel genes modulated in the thyroid of dogs treated with methimazole and propylthiouracil. J. Biol. Chem. 1996, 271, 28451–28457. [Google Scholar] [CrossRef]
  8. Sharova, L.V.; Sharov, A.A.; Nedorezov, T.; Piao, Y.; Shaik, N.; Ko, M.S. Database for mRNA half-life of 19 977 genes obtained by DNA microarray analysis of pluripotent and differentiating mouse embryonic stem cells. DNA Res. 2009, 16, 45–58. [Google Scholar] [CrossRef]
  9. Soubeyrand, S.; Martinuk, A.; Lau, P.; McPherson, R. TRIB1 Is Regulated Post-Transcriptionally by Proteasomal and Non-Proteasomal Pathways. PLoS ONE 2016, 11, e0152346. [Google Scholar] [CrossRef]
  10. Ruiz-Cantos, M.; Hutchison, C.E.; Shoulders, C.C. Musings from the Tribbles Research and Innovation Network. Cancers 2021, 13, 4517. [Google Scholar] [CrossRef] [PubMed]
  11. Fagerberg, L.; Hallström, B.M.; Oksvold, P.; Kampf, C.; Djureinovic, D.; Odeberg, J.; Habuka, M.; Tahmasebpoor, S.; Danielsson, A.; Edlund, K.; et al. Analysis of the human tissue-specific expression by genome-wide integration of transcriptomics and antibody-based proteomics. Mol. Cell Proteom. 2014, 13, 397–406. [Google Scholar] [CrossRef] [PubMed]
  12. Yokoyama, T.; Kanno, Y.; Yamazaki, Y.; Takahara, T.; Miyata, S.; Nakamura, T. Trib1 links the MEK1/ERK pathway in myeloid leukemogenesis. Blood 2010, 116, 2768–2775. [Google Scholar] [CrossRef] [PubMed]
  13. Kung, J.E.; Jura, N. The pseudokinase TRIB1 toggles an intramolecular switch to regulate COP1 nuclear export. EMBO J. 2019, 38, e99708. [Google Scholar] [CrossRef] [PubMed]
  14. Niespolo, C.; Johnston, J.M.; Deshmukh, S.R.; Satam, S.; Shologu, Z.; Villacanas, O.; Sudbery, I.M.; Wilson, H.L.; Kiss-Toth, E. Tribbles-1 Expression and Its Function to Control Inflammatory Cytokines, Including Interleukin-8 Levels are Regulated by miRNAs in Macrophages and Prostate Cancer Cells. Front. Immunol. 2020, 11, 574046. [Google Scholar] [CrossRef] [PubMed]
  15. Soubeyrand, S.; Lau, P.; McPherson, R. Regulation of TRIB1 abundance in hepatocyte models in response to proteasome inhibition. Sci. Rep. 2023, 13, 9320. [Google Scholar] [CrossRef] [PubMed]
  16. Mata, J.; Curado, S.; Ephrussi, A.; Rørth, P. Tribbles coordinates mitosis and morphogenesis in Drosophila by regulating string/CDC25 proteolysis. Cell 2000, 101, 511–522. [Google Scholar] [CrossRef] [PubMed]
  17. Grosshans, J.; Wieschaus, E. A genetic link between morphogenesis and cell division during formation of the ventral furrow in Drosophila. Cell 2000, 101, 523–531. [Google Scholar] [CrossRef] [PubMed]
  18. Seher, T.C.; Leptin, M. Tribbles, a cell-cycle brake that coordinates proliferation and morphogenesis during Drosophila gastrulation. Curr. Biol. 2000, 10, 623–629. [Google Scholar] [CrossRef]
  19. Kiss-Toth, E.; Wyllie, D.H.; Holland, K.; Marsden, L.; Jozsa, V.; Oxley, K.M.; Polgar, T.; Qwarnstrom, E.E.; Dower, S.K. Functional mapping and identification of novel regulators for the Toll/Interleukin-1 signalling network by transcription expression cloning. Cell Signal 2006, 18, 202–214. [Google Scholar] [CrossRef]
  20. Johnston, J.; Kiss-Toth, E. TRIB1 (tribbles pseudokinase 1). Atlas Genet. Cytogenet. Oncol. Haematol. 2016, 20, 106–114. [Google Scholar] [CrossRef]
  21. Foulkes, D.M.; Byrne, D.P.; Bailey, F.P.; Eyers, P.A. Tribbles pseudokinases: Novel targets for chemical biology and drug discovery? Biochem. Soc. Trans. 2015, 43, 1095–1103. [Google Scholar] [CrossRef] [PubMed]
  22. Murphy, J.M.; Nakatani, Y.; Jamieson, S.A.; Dai, W.; Lucet, I.S.; Mace, P.D. Molecular Mechanism of CCAAT-Enhancer Binding Protein Recruitment by the TRIB1 Pseudokinase. Structure 2015, 23, 2111–2121. [Google Scholar] [CrossRef] [PubMed]
  23. Jamieson, S.A.; Ruan, Z.; Burgess, A.E.; Curry, J.R.; McMillan, H.D.; Brewster, J.L.; Dunbier, A.K.; Axtman, A.D.; Kannan, N.; Mace, P.D. Substrate binding allosterically relieves autoinhibition of the pseudokinase TRIB1. Sci. Signal. 2018, 11, eaau0597. [Google Scholar] [CrossRef] [PubMed]
  24. Uljon, S.; Xu, X.; Durzynska, I.; Stein, S.; Adelmant, G.; Marto, J.A.; Pear, W.S.; Blacklow, S.C. Structural Basis for Substrate Selectivity of the E3 Ligase COP1. Structure 2016, 24, 687–696. [Google Scholar] [CrossRef] [PubMed]
  25. Goldberg, J.; Nairn, A.C.; Kuriyan, J. Structural basis for the autoinhibition of calcium/calmodulin-dependent protein kinase I. Cell 1996, 84, 875–887. [Google Scholar] [CrossRef] [PubMed]
  26. Eyers, P.A. TRIBBLES: A Twist in the Pseudokinase Tail. Structure 2015, 23, 1974–1976. [Google Scholar] [CrossRef] [PubMed]
  27. Yoshida, A.; Kato, J.Y.; Nakamae, I.; Yoneda-Kato, N. COP1 targets C/EBPα for degradation and induces acute myeloid leukemia via Trib1. Blood 2013, 122, 1750–1760. [Google Scholar] [CrossRef] [PubMed]
  28. Richmond, L.; Keeshan, K. Pseudokinases: A tribble-edged sword. FEBS J. 2020, 287, 4170–4182. [Google Scholar] [CrossRef]
  29. Rørth, P.; Szabo, K.; Texido, G. The level of C/EBP protein is critical for cell migration during Drosophila oogenesis and is tightly controlled by regulated degradation. Mol. Cell 2000, 6, 23–30. [Google Scholar] [CrossRef]
  30. Radomska, H.S.; Huettner, C.S.; Zhang, P.; Cheng, T.; Scadden, D.T.; Tenen, D.G. CCAAT/enhancer binding protein alpha is a regulatory switch sufficient for induction of granulocytic development from bipotential myeloid progenitors. Mol. Cell Biol. 1998, 18, 4301–4314. [Google Scholar] [CrossRef]
  31. Satoh, T.; Kidoya, H.; Naito, H.; Yamamoto, M.; Takemura, N.; Nakagawa, K.; Yoshioka, Y.; Morii, E.; Takakura, N.; Takeuchi, O.; et al. Critical role of Trib1 in differentiation of tissue-resident M2-like macrophages. Nature 2013, 495, 524–528. [Google Scholar] [CrossRef] [PubMed]
  32. Imajo, M.; Nishida, E. Human Tribbles homolog 1 functions as a negative regulator of retinoic acid receptor. Genes Cells 2010, 15, 1089–1097. [Google Scholar] [CrossRef] [PubMed]
  33. Das, R.; Sebo, Z.; Pence, L.; Dobens, L.L. Drosophila tribbles antagonizes insulin signaling-mediated growth and metabolism via interactions with Akt kinase. PLoS ONE 2014, 9, e109530. [Google Scholar] [CrossRef] [PubMed]
  34. Singh, K.; Han, C.; Fleming, J.L.; Becker, A.P.; McElroy, J.; Cui, T.; Johnson, B.; Kumar, A.; Sebastian, E.; Showalter, C.A.; et al. TRIB1 confers therapeutic resistance in GBM cells by activating the ERK and Akt pathways. Sci. Rep. 2023, 13, 12424. [Google Scholar] [CrossRef] [PubMed]
  35. Yu, J.M.; Sun, W.; Wang, Z.H.; Liang, X.; Hua, F.; Li, K.; Lv, X.X.; Zhang, X.W.; Liu, Y.Y.; Yu, J.J.; et al. TRIB3 supports breast cancer stemness by suppressing FOXO1 degradation and enhancing SOX2 transcription. Nat. Commun. 2019, 10, 5720. [Google Scholar] [CrossRef] [PubMed]
  36. Gendelman, R.; Xing, H.; Mirzoeva, O.K.; Sarde, P.; Curtis, C.; Feiler, H.S.; McDonagh, P.; Gray, J.W.; Khalil, I.; Korn, W.M. Bayesian Network Inference Modeling Identifies TRIB1 as a Novel Regulator of Cell-Cycle Progression and Survival in Cancer Cells. Cancer Res. 2017, 77, 1575–1585. [Google Scholar] [CrossRef] [PubMed]
  37. Kiss-Toth, E.; Bagstaff, S.M.; Sung, H.Y.; Jozsa, V.; Dempsey, C.; Caunt, J.C.; Oxley, K.M.; Wyllie, D.H.; Polgar, T.; Harte, M.; et al. Human tribbles, a protein family controlling mitogen-activated protein kinase cascades. J. Biol. Chem. 2004, 279, 42703–42708. [Google Scholar] [CrossRef] [PubMed]
  38. Sung, H.Y.; Guan, H.; Czibula, A.; King, A.R.; Eder, K.; Heath, E.; Suvarna, S.K.; Dower, S.K.; Wilson, A.G.; Francis, S.E.; et al. Human tribbles-1 controls proliferation and chemotaxis of smooth muscle cells via MAPK signaling pathways. J. Biol. Chem. 2007, 282, 18379–18387. [Google Scholar] [CrossRef]
  39. Arndt, L.; Dokas, J.; Gericke, M.; Kutzner, C.E.; Müller, S.; Jeromin, F.; Thiery, J.; Burkhardt, R. Tribbles homolog 1 deficiency modulates function and polarization of murine bone marrow-derived macrophages. J. Biol. Chem. 2018, 293, 11527–11536. [Google Scholar] [CrossRef]
  40. Ishizuka, Y.; Nakayama, K.; Ogawa, A.; Makishima, S.; Boonvisut, S.; Hirao, A.; Iwasaki, Y.; Yada, T.; Yanagisawa, Y.; Miyashita, H.; et al. TRIB1 downregulates hepatic lipogenesis and glycogenesis via multiple molecular interactions. J. Mol. Endocrinol. 2014, 52, 145–158. [Google Scholar] [CrossRef]
  41. Soubeyrand, S.; Martinuk, A.; McPherson, R. TRIB1 is a positive regulator of hepatocyte nuclear factor 4-alpha. Sci. Rep. 2017, 7, 5574. [Google Scholar] [CrossRef] [PubMed]
  42. Makishima, S.; Boonvisut, S.; Ishizuka, Y.; Watanabe, K.; Nakayama, K.; Iwamoto, S. Sin3A-associated protein, 18 kDa, a novel binding partner of TRIB1, regulates MTTP expression. J. Lipid Res. 2015, 56, 1145–1152. [Google Scholar] [CrossRef] [PubMed]
  43. Sun, X.; Wang, S.; Miao, X.; Zeng, S.; Guo, Y.; Zhou, A.; Chen, Y.; Lv, F.; Fan, Z.; Wang, Y.; et al. TRIB1 regulates liver regeneration by antagonizing the NRF2-mediated antioxidant response. Cell Death Dis. 2023, 14, 372. [Google Scholar] [CrossRef]
  44. Ostertag, A.; Jones, A.; Rose, A.J.; Liebert, M.; Kleinsorg, S.; Reimann, A.; Vegiopoulos, A.; Berriel Diaz, M.; Strzoda, D.; Yamamoto, M.; et al. Control of adipose tissue inflammation through TRB1. Diabetes 2010, 59, 1991–2000. [Google Scholar] [CrossRef] [PubMed]
  45. Dugast, E.; Kiss-Toth, E.; Docherty, L.; Danger, R.; Chesneau, M.; Pichard, V.; Judor, J.P.; Pettré, S.; Conchon, S.; Soulillou, J.P.; et al. Identification of tribbles-1 as a novel binding partner of Foxp3 in regulatory T cells. J. Biol. Chem. 2013, 288, 10051–10060. [Google Scholar] [CrossRef]
  46. McMillan, H.D.; Papachristou, E.K.; Hazlett, J.; Omarjee, S.; Carroll, J.S.; Black, M.A.; Mace, P.D.; Dunbier, A.K. TRIB1 modulates transcriptional programming in breast cancer cells to regulate cell proliferation. bioRxiv 2023. [Google Scholar] [CrossRef]
  47. Miyajima, C.; Inoue, Y.; Hayashi, H. Pseudokinase tribbles 1 (TRB1) negatively regulates tumor-suppressor activity of p53 through p53 deacetylation. Biol. Pharm. Bull. 2015, 38, 618–624. [Google Scholar] [CrossRef] [PubMed]
  48. Simoni, L.; Delgado, V.; Ruer-Laventie, J.; Bouis, D.; Soley, A.; Heyer, V.; Robert, I.; Gies, V.; Martin, T.; Korganow, A.S.; et al. Trib1 Is Overexpressed in Systemic Lupus Erythematosus, While It Regulates Immunoglobulin Production in Murine B Cells. Front. Immunol. 2018, 9, 373. [Google Scholar] [CrossRef] [PubMed]
  49. Rome, K.S.; Stein, S.J.; Kurachi, M.; Petrovic, J.; Schwartz, G.W.; Mack, E.A.; Uljon, S.; Wu, W.W.; DeHart, A.G.; McClory, S.E.; et al. Trib1 regulates T cell differentiation during chronic infection by restraining the effector program. J. Exp. Med. 2020, 217, e20190888. [Google Scholar] [CrossRef]
  50. Kathiresan, S.; Melander, O.; Guiducci, C.; Surti, A.; Burtt, N.P.; Rieder, M.J.; Cooper, G.M.; Roos, C.; Voight, B.F.; Havulinna, A.S.; et al. Six new loci associated with blood low-density lipoprotein cholesterol, high-density lipoprotein cholesterol or triglycerides in humans. Nat. Genet. 2008, 40, 189–197. [Google Scholar] [CrossRef]
  51. Burkhardt, R.; Toh, S.A.; Lagor, W.R.; Birkeland, A.; Levin, M.; Li, X.; Robblee, M.; Fedorov, V.D.; Yamamoto, M.; Satoh, T.; et al. Trib1 is a lipid- and myocardial infarction-associated gene that regulates hepatic lipogenesis and VLDL production in mice. J. Clin. Investig. 2010, 120, 4410–4414. [Google Scholar] [CrossRef] [PubMed]
  52. Bauer, R.C.; Sasaki, M.; Cohen, D.M.; Cui, J.; Smith, M.A.; Yenilmez, B.O.; Steger, D.J.; Rader, D.J. Tribbles-1 regulates hepatic lipogenesis through posttranscriptional regulation of C/EBPα. J. Clin. Investig. 2015, 125, 3809–3818. [Google Scholar] [CrossRef] [PubMed]
  53. Quiroz-Figueroa, K.; Vitali, C.; Conlon, D.M.; Millar, J.S.; Tobias, J.W.; Bauer, R.C.; Hand, N.J.; Rader, D.J. TRIB1 regulates LDL metabolism through CEBPα-mediated effects on the LDL receptor in hepatocytes. J. Clin. Investig. 2021, 131, e146775. [Google Scholar] [CrossRef] [PubMed]
  54. Ha, E.E.; Quartuccia, G.I.; Ling, R.; Xue, C.; Karikari, R.A.; Hernandez-Ono, A.; Hu, K.Y.; Matias, C.V.; Imam, R.; Cui, J.; et al. Adipocyte-specific tribbles pseudokinase 1 regulates plasma adiponectin and plasma lipids in mice. Mol. Metab. 2022, 56, 101412. [Google Scholar] [CrossRef] [PubMed]
  55. Chambers, J.C.; Zhang, W.; Sehmi, J.; Li, X.; Wass, M.N.; Van der Harst, P.; Holm, H.; Sanna, S.; Kavousi, M.; Baumeister, S.E.; et al. Genome-wide association study identifies loci influencing concentrations of liver enzymes in plasma. Nat. Genet. 2011, 43, 1131–1138. [Google Scholar] [CrossRef] [PubMed]
  56. Bhairavabhotla, R.; Kim, Y.C.; Glass, D.D.; Escobar, T.M.; Patel, M.C.; Zahr, R.; Nguyen, C.K.; Kilaru, G.K.; Muljo, S.A.; Shevach, E.M. Transcriptome profiling of human FoxP3+ regulatory T cells. Hum. Immunol. 2016, 77, 201–213. [Google Scholar] [CrossRef] [PubMed]
  57. Miyajima, C.; Itoh, Y.; Inoue, Y.; Hayashi, H. Positive Regulation of Interleukin-2 Expression by a Pseudokinase, Tribbles 1, in Activated T Cells. Biol. Pharm. Bull. 2015, 38, 1126–1133. [Google Scholar] [CrossRef] [PubMed]
  58. Yamamoto, M.; Uematsu, S.; Okamoto, T.; Matsuura, Y.; Sato, S.; Kumar, H.; Satoh, T.; Saitoh, T.; Takeda, K.; Ishii, K.J.; et al. Enhanced TLR-mediated NF-IL6 dependent gene expression by Trib1 deficiency. J. Exp. Med. 2007, 204, 2233–2239. [Google Scholar] [CrossRef] [PubMed]
  59. Liu, Y.H.; Tan, K.A.; Morrison, I.W.; Lamb, J.R.; Argyle, D.J. Macrophage migration is controlled by Tribbles 1 through the interaction between C/EBPβ and TNF-α. Vet. Immunol. Immunopathol. 2013, 155, 67–75. [Google Scholar] [CrossRef]
  60. Ashton-Chess, J.; Giral, M.; Mengel, M.; Renaudin, K.; Foucher, Y.; Gwinner, W.; Braud, C.; Dugast, E.; Quillard, T.; Thebault, P.; et al. Tribbles-1 as a novel biomarker of chronic antibody-mediated rejection. J. Am. Soc. Nephrol. 2008, 19, 1116–1127. [Google Scholar] [CrossRef]
  61. Röthlisberger, B.; Heizmann, M.; Bargetzi, M.J.; Huber, A.R. TRIB1 overexpression in acute myeloid leukemia. Cancer Genet. Cytogenet. 2007, 176, 58–60. [Google Scholar] [CrossRef]
  62. Döhner, H. Implication of the molecular characterization of acute myeloid leukemia. Hematol. Am. Soc. Hematol. Educ. Program. 2007, 2007, 412–419. [Google Scholar] [CrossRef] [PubMed]
  63. Yoshino, S.; Yokoyama, T.; Sunami, Y.; Takahara, T.; Nakamura, A.; Yamazaki, Y.; Tsutsumi, S.; Aburatani, H.; Nakamura, T. Trib1 promotes acute myeloid leukemia progression by modulating the transcriptional programs of Hoxa9. Blood 2021, 137, 75–88. [Google Scholar] [CrossRef]
  64. Pineault, N.; Helgason, C.D.; Lawrence, H.J.; Humphries, R.K. Differential expression of Hox, Meis1, and Pbx1 genes in primitive cells throughout murine hematopoietic ontogeny. Exp. Hematol. 2002, 30, 49–57. [Google Scholar] [CrossRef]
  65. Jin, G.; Yamazaki, Y.; Takuwa, M.; Takahara, T.; Kaneko, K.; Kuwata, T.; Miyata, S.; Nakamura, T. Trib1 and Evi1 cooperate with Hoxa and Meis1 in myeloid leukemogenesis. Blood 2007, 109, 3998–4005. [Google Scholar] [CrossRef] [PubMed]
  66. Kroon, E.; Krosl, J.; Thorsteinsdottir, U.; Baban, S.; Buchberg, A.M.; Sauvageau, G. Hoxa9 transforms primary bone marrow cells through specific collaboration with Meis1a but not Pbx1b. EMBO J. 1998, 17, 3714–3725. [Google Scholar] [CrossRef] [PubMed]
  67. Lawrence, H.J.; Rozenfeld, S.; Cruz, C.; Matsukuma, K.; Kwong, A.; Kömüves, L.; Buchberg, A.M.; Largman, C. Frequent co-expression of the HOXA9 and MEIS1 homeobox genes in human myeloid leukemias. Leukemia 1999, 13, 1993–1999. [Google Scholar] [CrossRef]
  68. Mohr, S.; Doebele, C.; Comoglio, F.; Berg, T.; Beck, J.; Bohnenberger, H.; Alexe, G.; Corso, J.; Ströbel, P.; Wachter, A.; et al. Hoxa9 and Meis1 Cooperatively Induce Addiction to Syk Signaling by Suppressing miR-146a in Acute Myeloid Leukemia. Cancer Cell 2017, 31, 549–562.e11. [Google Scholar] [CrossRef]
  69. Shahrouzi, P.; Astobiza, I.; Cortazar, A.R.; Torrano, V.; Macchia, A.; Flores, J.M.; Niespolo, C.; Mendizabal, I.; Caloto, R.; Ercilla, A.; et al. Genomic and Functional Regulation of TRIB1 Contributes to Prostate Cancer Pathogenesis. Cancers 2020, 12, 2593. [Google Scholar] [CrossRef]
  70. Mashima, T.; Soma-Nagae, T.; Migita, T.; Kinoshita, R.; Iwamoto, A.; Yuasa, T.; Yonese, J.; Ishikawa, Y.; Seimiya, H. TRIB1 supports prostate tumorigenesis and tumor-propagating cell survival by regulation of endoplasmic reticulum chaperone expression. Cancer Res. 2014, 74, 4888–4897. [Google Scholar] [CrossRef]
  71. Liu, Z.Z.; Han, Z.D.; Liang, Y.K.; Chen, J.X.; Wan, S.; Zhuo, Y.J.; Cai, Z.D.; Deng, Y.L.; Lin, Z.Y.; Mo, R.J.; et al. TRIB1 induces macrophages to M2 phenotype by inhibiting IKB-zeta in prostate cancer. Cell Signal 2019, 59, 152–162. [Google Scholar] [CrossRef] [PubMed]
  72. Papachristou, E.K.; Kishore, K.; Holding, A.N.; Harvey, K.; Roumeliotis, T.I.; Chilamakuri, C.S.R.; Omarjee, S.; Chia, K.M.; Swarbrick, A.; Lim, E.; et al. A quantitative mass spectrometry-based approach to monitor the dynamics of endogenous chromatin-associated protein complexes. Nat. Commun. 2018, 9, 2311. [Google Scholar] [CrossRef] [PubMed]
  73. Kim, T.; Johnston, J.; Castillo-Lluva, S.; Cimas, F.J.; Hamby, S.; Gonzalez-Moreno, S.; Villarejo-Campos, P.; Goodall, A.H.; Velasco, G.; Ocana, A.; et al. TRIB1 regulates tumor growth via controlling tumor-associated macrophage phenotypes and is associated with breast cancer survival and treatment response. Theranostics 2022, 12, 3584–3600. [Google Scholar] [CrossRef] [PubMed]
  74. Camps, J.; Nguyen, Q.T.; Padilla-Nash, H.M.; Knutsen, T.; McNeil, N.E.; Wangsa, D.; Hummon, A.B.; Grade, M.; Ried, T.; Difilippantonio, M.J. Integrative genomics reveals mechanisms of copy number alterations responsible for transcriptional deregulation in colorectal cancer. Genes. Chromosomes Cancer 2009, 48, 1002–1017. [Google Scholar] [CrossRef] [PubMed]
  75. Briffa, R.; Um, I.; Faratian, D.; Zhou, Y.; Turnbull, A.K.; Langdon, S.P.; Harrison, D.J. Multi-Scale Genomic, Transcriptomic and Proteomic Analysis of Colorectal Cancer Cell Lines to Identify Novel Biomarkers. PLoS ONE 2015, 10, e0144708. [Google Scholar] [CrossRef] [PubMed]
  76. Wang, Y.; Wu, N.; Pang, B.; Tong, D.; Sun, D.; Sun, H.; Zhang, C.; Sun, W.; Meng, X.; Bai, J.; et al. TRIB1 promotes colorectal cancer cell migration and invasion through activation MMP-2 via FAK/Src and ERK pathways. Oncotarget 2017, 8, 47931–47942. [Google Scholar] [CrossRef] [PubMed]
  77. Lu, X.X.; Hu, J.J.; Fang, Y.; Wang, Z.T.; Xie, J.J.; Zhan, Q.; Deng, X.X.; Chen, H.; Jin, J.B.; Peng, C.H.; et al. A case-control study indicates that the TRIB1 gene is associated with pancreatic cancer. Genet. Mol. Res. 2014, 13, 6142–6147. [Google Scholar] [CrossRef] [PubMed]
  78. Chen, H.; Li, M.; Sanchez, E.; Soof, C.M.; Bujarski, S.; Ng, N.; Cao, J.; Hekmati, T.; Zahab, B.; Nosrati, J.D.; et al. JAK1/2 pathway inhibition suppresses M2 polarization and overcomes resistance of myeloma to lenalidomide by reducing TRIB1, MUC1, CD44, CXCL12, and CXCR4 expression. Br. J. Haematol. 2020, 188, 283–294. [Google Scholar] [CrossRef] [PubMed]
  79. Ye, Y.; Wang, G.; Zhuang, J.; He, S.; Song, Y.; Ni, J.; Xia, W.; Wang, J. The Oncogenic Role of Tribbles 1 in Hepatocellular Carcinoma Is Mediated by a Feedback Loop Involving microRNA-23a and p53. Front. Physiol. 2017, 8, 789. [Google Scholar] [CrossRef]
  80. Tang, B.; Wu, W.; Zhang, Q.; Sun, Y.; Cui, Y.; Wu, F.; Wei, X.; Qi, G.; Liang, X.; Tang, F.; et al. Inhibition of tribbles protein-1 attenuates radioresistance in human glioma cells. Sci. Rep. 2015, 5, 15961. [Google Scholar] [CrossRef]
  81. Bauer, R.C.; Yenilmez, B.O.; Rader, D.J. Tribbles-1: A novel regulator of hepatic lipid metabolism in humans. Biochem. Soc. Trans. 2015, 43, 1079–1084. [Google Scholar] [CrossRef] [PubMed]
  82. Arndt, L.; Hernandez-Resendiz, I.; Moos, D.; Dokas, J.; Müller, S.; Jeromin, F.; Wagner, R.; Ceglarek, U.; Heid, I.M.; Höring, M.; et al. Deficiency Promotes Hyperlipidemia, Inflammation, and Atherosclerosis in LDL Receptor Knockout Mice. Arterioscler. Thromb. Vasc. Biol. 2023, 43, 979–994. [Google Scholar] [CrossRef]
  83. Johnston, J.M.; Angyal, A.; Bauer, R.C.; Hamby, S.; Suvarna, S.K.; Baidžajevas, K.; Hegedus, Z.; Dear, T.N.; Turner, M.; Wilson, H.L.; et al. Myeloid Tribbles 1 induces early atherosclerosis via enhanced foam cell expansion. Sci. Adv. 2019, 5, eaax9183. [Google Scholar] [CrossRef] [PubMed]
  84. Jiang, J.; Chen, X.; Li, C.; Du, X.; Zhou, H. Polymorphisms of TRIB1 genes for coronary artery disease and stroke risk: A systematic review and meta-analysis. Gene 2023, 880, 147613. [Google Scholar] [CrossRef] [PubMed]
  85. Kitamoto, A.; Kitamoto, T.; Nakamura, T.; Ogawa, Y.; Yoneda, M.; Hyogo, H.; Ochi, H.; Mizusawa, S.; Ueno, T.; Nakao, K.; et al. Association of polymorphisms in GCKR and TRIB1 with nonalcoholic fatty liver disease and metabolic syndrome traits. Endocr. J. 2014, 61, 683–689. [Google Scholar] [CrossRef] [PubMed]
  86. Liu, Q.; Xue, F.; Meng, J.; Liu, S.S.; Chen, L.Z.; Gao, H.; Geng, N.; Jin, W.W.; Xin, Y.N.; Xuan, S.Y. TRIB1 rs17321515 and rs2954029 gene polymorphisms increase the risk of non-alcoholic fatty liver disease in Chinese Han population. Lipids Health Dis. 2019, 18, 61. [Google Scholar] [CrossRef] [PubMed]
  87. Soubeyrand, S.; Martinuk, A.; Naing, T.; Lau, P.; McPherson, R. Role of Tribbles Pseudokinase 1 (TRIB1) in human hepatocyte metabolism. Biochim. Biophys. Acta 2016, 1862, 223–232. [Google Scholar] [CrossRef] [PubMed]
  88. Nagiec, M.M.; Skepner, A.P.; Negri, J.; Eichhorn, M.; Kuperwasser, N.; Comer, E.; Muncipinto, G.; Subramanian, A.; Clish, C.; Musunuru, K.; et al. Modulators of hepatic lipoprotein metabolism identified in a search for small-molecule inducers of tribbles pseudokinase 1 expression. PLoS ONE 2015, 10, e0120295. [Google Scholar] [CrossRef] [PubMed]
  89. Cao, A.; Wu, M.; Li, H.; Liu, J. Janus kinase activation by cytokine oncostatin M decreases PCSK9 expression in liver cells. J. Lipid Res. 2011, 52, 518–530. [Google Scholar] [CrossRef]
  90. Nagiec, M.M.; Duvall, J.R.; Skepner, A.P.; Howe, E.A.; Bastien, J.; Comer, E.; Marie, J.C.; Johnston, S.E.; Negri, J.; Eichhorn, M.; et al. Novel tricyclic glycal-based. MedChemComm 2018, 9, 1831–1842. [Google Scholar] [CrossRef]
  91. Zhang, X.; Zhang, B.; Zhang, C.; Sun, G.; Sun, X. Trib1 deficiency causes brown adipose respiratory chain depletion and mitochondrial disorder. Cell Death Dis. 2021, 12, 1098. [Google Scholar] [CrossRef] [PubMed]
  92. De Marco, C.; Laudanna, C.; Rinaldo, N.; Oliveira, D.M.; Ravo, M.; Weisz, A.; Ceccarelli, M.; Caira, E.; Rizzuto, A.; Zoppoli, P.; et al. Specific gene expression signatures induced by the multiple oncogenic alterations that occur within the PTEN/PI3K/AKT pathway in lung cancer. PLoS ONE 2017, 12, e0178865. [Google Scholar] [CrossRef]
  93. Wang, L.; Liu, X.; Ren, Y.; Zhang, J.; Chen, J.; Zhou, W.; Guo, W.; Wang, X.; Chen, H.; Li, M.; et al. Cisplatin-enriching cancer stem cells confer multidrug resistance in non-small cell lung cancer via enhancing TRIB1/HDAC activity. Cell Death Dis. 2017, 8, e2746. [Google Scholar] [CrossRef]
  94. Jurj, A.; Pop, L.A.; Zanoaga, O.; Ciocan-Cârtiţă, C.A.; Cojocneanu, R.; Moldovan, C.; Raduly, L.; Pop-Bica, C.; Trif, M.; Irimie, A.; et al. New Insights in Gene Expression Alteration as Effect of Paclitaxel Drug Resistance in Triple Negative Breast Cancer Cells. Cell Physiol. Biochem. 2020, 54, 648–664. [Google Scholar] [CrossRef] [PubMed]
  95. Keeshan, K.; Vieugué, P.; Chaudhury, S.; Rishi, L.; Gaillard, C.; Liang, L.; Garcia, E.; Nakamura, T.; Omidvar, N.; Kogan, S.C. Co-operative leukemogenesis in acute myeloid leukemia and acute promyelocytic leukemia reveals C/EBPα as a common target of TRIB1 and PML/RARA. Haematologica 2016, 101, 1228–1236. [Google Scholar] [CrossRef]
  96. McClory, S.E.; Bardhan, O.; Rome, K.S.; Giles, J.R.; Baxter, A.E.; Xu, L.; Gimotty, P.A.; Faryabi, R.B.; Wherry, E.J.; Pear, W.S.; et al. The pseudokinase Trib1 regulates the transition of exhausted T cells to a KLR. Cell Rep. 2023, 42, 112905. [Google Scholar] [CrossRef] [PubMed]
  97. Shang, S.; Yang, Y.W.; Chen, F.; Yu, L.; Shen, S.H.; Li, K.; Cui, B.; Lv, X.X.; Zhang, C.; Yang, C.; et al. TRIB3 reduces CD8. Sci. Transl. Med. 2022, 14, eabf0992. [Google Scholar] [CrossRef]
  98. Lin, Z.Y.; Huang, Y.Q.; Zhang, Y.Q.; Han, Z.D.; He, H.C.; Ling, X.H.; Fu, X.; Dai, Q.S.; Cai, C.; Chen, J.H.; et al. MicroRNA-224 inhibits progression of human prostate cancer by downregulating TRIB1. Int. J. Cancer 2014, 135, 541–550. [Google Scholar] [CrossRef] [PubMed]
  99. Tan, X.; Song, X.; Fan, B.; Li, M.; Zhang, A.; Pei, L. Exosomal circRNA Scm-like with four malignant brain tumor domains 2 (circ-SFMBT2) enhances the docetaxel resistance of prostate cancer via the microRNA-136-5p/tribbles homolog 1 pathway. Anticancer Drugs 2022, 33, 871–882. [Google Scholar] [CrossRef]
  100. Györffy, B.; Surowiak, P.; Kiesslich, O.; Denkert, C.; Schäfer, R.; Dietel, M.; Lage, H. Gene expression profiling of 30 cancer cell lines predicts resistance towards 11 anticancer drugs at clinically achieved concentrations. Int. J. Cancer 2006, 118, 1699–1712. [Google Scholar] [CrossRef]
  101. Xie, X.; Lee, J.; Fuson, J.A.; Liu, H.; Iwase, T.; Yun, K.; Margain, C.; Tripathy, D.; Ueno, N.T. Identification of Kinase Targets for Enhancing the Antitumor Activity of Eribulin in Triple-Negative Breast Cell Lines. Biomedicines 2023, 11, 735. [Google Scholar] [CrossRef]
  102. Allen, W.L.; Coyle, V.M.; Jithesh, P.V.; Proutski, I.; Stevenson, L.; Fenning, C.; Longley, D.B.; Wilson, R.H.; Gordon, M.; Lenz, H.J.; et al. Clinical determinants of response to irinotecan-based therapy derived from cell line models. Clin. Cancer Res. 2008, 14, 6647–6655. [Google Scholar] [CrossRef]
  103. Shawky, A.M.; Almalki, F.A.; Abdalla, A.N.; Abdelazeem, A.H.; Gouda, A.M. A Comprehensive Overview of Globally Approved JAK Inhibitors. Pharmaceutics 2022, 14, 1001. [Google Scholar] [CrossRef]
  104. Mascarenhas, J.; Hoffman, R. Ruxolitinib: The first FDA approved therapy for the treatment of myelofibrosis. Clin. Cancer Res. 2012, 18, 3008–3014. [Google Scholar] [CrossRef] [PubMed]
  105. Foulkes, D.M.; Byrne, D.P.; Yeung, W.; Shrestha, S.; Bailey, F.P.; Ferries, S.; Eyers, C.E.; Keeshan, K.; Wells, C.; Drewry, D.H.; et al. Covalent inhibitors of EGFR family protein kinases induce degradation of human Tribbles 2 (TRIB2) pseudokinase in cancer cells. Sci. Signal. 2018, 11, eaat7951. [Google Scholar] [CrossRef]
  106. Lim, S.M.; Westover, K.D.; Ficarro, S.B.; Harrison, R.A.; Choi, H.G.; Pacold, M.E.; Carrasco, M.; Hunter, J.; Kim, N.D.; Xie, T.; et al. Therapeutic targeting of oncogenic K-Ras by a covalent catalytic site inhibitor. Angew. Chem. Int. Ed. Engl. 2014, 53, 199–204. [Google Scholar] [CrossRef]
  107. Lai, A.C.; Crews, C.M. Induced protein degradation: An emerging drug discovery paradigm. Nat. Rev. Drug Discov. 2017, 16, 101–114. [Google Scholar] [CrossRef]
  108. Neklesa, T.K.; Crews, C.M. Chemical biology: Greasy tags for protein removal. Nature 2012, 487, 308–309. [Google Scholar] [CrossRef]
  109. Alabi, S.B. Targeting Oncogenic Kinases and Pseudokinases with Proteolysis Targeting Chimeras; Yale University: New Haven, CT, USA, 2021. [Google Scholar]
  110. Xie, T.; Lim, S.M.; Westover, K.D.; Dodge, M.E.; Ercan, D.; Ficarro, S.B.; Udayakumar, D.; Gurbani, D.; Tae, H.S.; Riddle, S.M.; et al. Pharmacological targeting of the pseudokinase Her3. Nat. Chem. Biol. 2014, 10, 1006–1012. [Google Scholar] [CrossRef]
  111. Degorce, S.L.; Tavana, O.; Banks, E.; Crafter, C.; Gingipalli, L.; Kouvchinov, D.; Mao, Y.; Pachl, F.; Solanki, A.; Valge-Archer, V.; et al. Discovery of Proteolysis-Targeting Chimera Molecules that Selectively Degrade the IRAK3 Pseudokinase. J. Med. Chem. 2020, 63, 10460–10473. [Google Scholar] [CrossRef]
  112. Rathje, O.H.; Perryman, L.; Payne, R.J.; Hamprecht, D.W. PROTACs Targeting MLKL Protect Cells from Necroptosis. J. Med. Chem. 2023, 66, 11216–11236. [Google Scholar] [CrossRef] [PubMed]
  113. Sanchez-Vega, F.; Mina, M.; Armenia, J.; Chatila, W.K.; Luna, A.; La, K.C.; Dimitriadoy, S.; Liu, D.L.; Kantheti, H.S.; Saghafinia, S.; et al. Oncogenic Signaling Pathways in The Cancer Genome Atlas. Cell 2018, 173, 321–337.e10. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Overview of TRIB1 structure and functions. TRIB1 has three domains: An N-terminal domain, a pseudokinase domain, and a C-terminal domain. All three domains bind to distinct substrates and interacting proteins to carry out different functions. The N-terminal domain contains a NLS (nuclear localization signal/sequence) and a typical PEST domain. It is shown to bind MALT1 and regulate TCR (T cell receptor) signaling. The pseudokinase domain is the longest domain and acts as a scaffold and sequester a variety of proteins. It is known to bind NF-кB, Foxp3, HDAC1, HNF4A, Akt, and MKK4 and influence processes such as inflammation, cell cycle regulation, p53 deacetylation, lipid metabolism, Treg regulation, cell survival, and proliferation. The pseudokinase domain also binds substrates for proteasomal degradation such as C/EBPα/β, MLXIPL (ChREBP), FERMT2, RAR/RXRα, and CD72. Degradation of these substrates affects processes such as myeloid cell differentiation, lipid metabolism, adipocyte maintenance, transcriptional downregulation, and autoimmunity. The pseudokinase domain also contains an SLE sequence as a substitute for the DFG motif that is responsible for binding ATP in a canonical kinase. The C-terminal domain binds MEK1/2 at ILLHPW sequence and activates downstream MAPK signaling to induce cell proliferation. COP1 E3 ligase also binds at the DQIVPE sequence in the C-terminal domain to result in degradation of TRIB1 substrates through the ubiquitin proteasome system. CUL4A/B is another ubiquitin ligase that is shown to bind TRIB1. Nrf2 also binds to TRIB1 through the C-terminal domain, and this interaction has been implicated in liver regeneration.
Figure 1. Overview of TRIB1 structure and functions. TRIB1 has three domains: An N-terminal domain, a pseudokinase domain, and a C-terminal domain. All three domains bind to distinct substrates and interacting proteins to carry out different functions. The N-terminal domain contains a NLS (nuclear localization signal/sequence) and a typical PEST domain. It is shown to bind MALT1 and regulate TCR (T cell receptor) signaling. The pseudokinase domain is the longest domain and acts as a scaffold and sequester a variety of proteins. It is known to bind NF-кB, Foxp3, HDAC1, HNF4A, Akt, and MKK4 and influence processes such as inflammation, cell cycle regulation, p53 deacetylation, lipid metabolism, Treg regulation, cell survival, and proliferation. The pseudokinase domain also binds substrates for proteasomal degradation such as C/EBPα/β, MLXIPL (ChREBP), FERMT2, RAR/RXRα, and CD72. Degradation of these substrates affects processes such as myeloid cell differentiation, lipid metabolism, adipocyte maintenance, transcriptional downregulation, and autoimmunity. The pseudokinase domain also contains an SLE sequence as a substitute for the DFG motif that is responsible for binding ATP in a canonical kinase. The C-terminal domain binds MEK1/2 at ILLHPW sequence and activates downstream MAPK signaling to induce cell proliferation. COP1 E3 ligase also binds at the DQIVPE sequence in the C-terminal domain to result in degradation of TRIB1 substrates through the ubiquitin proteasome system. CUL4A/B is another ubiquitin ligase that is shown to bind TRIB1. Nrf2 also binds to TRIB1 through the C-terminal domain, and this interaction has been implicated in liver regeneration.
Cancers 16 01889 g001
Table 2. TRIB1 in cancer and treatment resistance.
Table 2. TRIB1 in cancer and treatment resistance.
CancerBiological Significance of TRIB1ReferencesCancer
Treatment
Role of TRIB1 in Treatment
Resistance
References
Acute myeloid leukemiaTRIB1 overexpression drives Hoxa9-induced leukemogenesis by decreasing C/EBPα to induce the enhancer programs at Erg and Spns2 loci. TRIB1 interacts with MEK1 to activate the MAPK pathway and enhances self-renewal of malignant bone marrow cells.[12,63]All-trans retinoic acidTRIB1 overexpression causes resistance to ATRA treatment in acute promyelocytic leukemia through downregulation of C/EBPα.[95]
Cytarabine
Daunorubicin
Idarubicin
None reported
Prostate cancerTRIB1 and cMYC are co-amplified in prostate cancer. TRIB1 promotes secretion of CXCL and IL-8 by inhibiting IкB-ζ expression and induces tumor growth. TRIB1 upregulates GRP78 to promote the occurrence and survival of prostate tumor cells. MiR-224, which targets TRIB1, is downregulated in prostate cancer.[70,71,98]DocetaxelIncreased expression of TRIB1 promotes resistance to docetaxel.[99]
MitoxantroneIncreased TRIB1 gene expression is associated with resistance to mitoxantrone.[100]
Estramustine
Carboplatin
None reported
Breast cancerTRIB1 and cMYC are co-amplified in breast cancer patients. TRIB1 regulates both G1/S and G2/M transition in breast cancer cells. TRIB1 interacts with β-catenin and its co-factor FERMT2 and may regulate β-catenin activity. TRIB1 also interacts with ER co-factor ZBTB7A and could influence ER-associated transcription. TRIB1 reduces DR5 and TRAIL-induced apoptosis by elevating NF-кB signaling. TRIB1 negatively regulates the anti-tumor cytokine IL-15 in tumor associated macrophages to decrease T-cell infiltration and anti-tumor responses.[36,46,73]PaclitaxelLong-term paclitaxel treatment in breast cancer cell lines leads to induction of TRIB1 gene and paclitaxel resistance.[94]
EribulinInhibition of TRIB1 enhanced the anti-proliferation effect of eribulin in triple-negative breast cancer cells.[101]
Doxorubicin
Epirubicin
5-fluorouracil (5-FU)

Capecitabine
Cyclophosphamide
Carboplatin
Ixabepilone
Vinorelbine
Gemcitabine
None reported
Colorectal cancerTRIB1 and cMYC are co-amplified in cells and in patients. TRIB1 promotes migration and invasion of CRC cells through the activation of FAK/Src and the ERK pathway to upregulate MMP-2 expression.[74,76]IrinotecanTreatment with the active metabolite of irinotecan, SN38, acutely induces TRIB1 gene expression in colon cancer cells. Increased TRIB1 gene expression is associated with resistance to irinotecan.[102]
5-FU
Capecitabine
Oxaliplatin
Trifluridine and tipiracil
None reported
Pancreatic cancerA single nucleotide polymorphism in the TRIB1 gene (rs2980879) is associated with pancreatic cancer in the Chinese Han population.[77]Gemcitabine
5-FU
Oxaliplatin
Albumin-bound paclitaxel (Abrxane)
Capecitabine
Cisplatin
Irinotecan
None reported
Multiple myelomaTRIB1 gene expression is higher in bone marrow mononuclear cells of multiple myeloma patients with progressive disease compared to those in remission. Patients with progressive disease have increased percentage of M2 macrophages and suggests a role of TRIB1 in M2 macrophage polarization in multiple myeloma potentially via the JAK/STAT pathway.[78]Cyclophosphamide
Etoposide
Doxorubicin
Melphalan
Bendamustine
None reported
Hepatocellular carcinomaTRIB1 is upregulated in hepatocellular carcinoma cells and tissues. TRIB1 promotes hepatocellular carcinoma tumorigenesis and invasiveness through a feedback loop involving miR-23a and p53.[79]Sorafenib
Gemcitabine
Oxaliplatin
Cisplatin
Doxorubicin
5-FU
Capecitabine
Mitoxantrone
None reported
GliomaTRIB1 causes upregulation of ERK and Akt pathways. TRIB1 inhibits activity of p53 in glioma through COP1 and HDAC1, leading to treatment resistance.[34,80]RadiationTRIB1 mRNA and protein levels increase after radiation treatment in glioma cells causing a decrease in treatment-induced cell death.[34,80]
TemozolomideTRIB1 mRNA and protein levels increase after temozolomide treatment in glioma cells causing a decrease in treatment-induced cell death.[34,80]
Non-small cell lung cancerTRIB1 interacts with HDAC1 to deacetylate and inactivate p53. TRIB1 participates in the abnormal activation of the PI3K/Akt pathway through regulation by mutant PIK3CA.[92,93]CisplatinCisplatin treatment resulted in C/EBP-β-dependent increasing of TRIB1, which forms a complex with HDAC1 to downregulate p53 and induce resistance to cisplatin.[93]
VinorelbineCisplatin pre-treatment of lung cancer cells resulted in increased resistance to vinorelbine.[93]
Carboplatin
Paclitaxel
Albumin-bound paclitaxel (Abraxane)
Docetaxel
Gemcitabine
Etoposide
Pemetrexed
None 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

Singh, K.; Showalter, C.A.; Manring, H.R.; Haque, S.J.; Chakravarti, A. “Oh, Dear We Are in Tribble”: An Overview of the Oncogenic Functions of Tribbles 1. Cancers 2024, 16, 1889. https://doi.org/10.3390/cancers16101889

AMA Style

Singh K, Showalter CA, Manring HR, Haque SJ, Chakravarti A. “Oh, Dear We Are in Tribble”: An Overview of the Oncogenic Functions of Tribbles 1. Cancers. 2024; 16(10):1889. https://doi.org/10.3390/cancers16101889

Chicago/Turabian Style

Singh, Karnika, Christian A. Showalter, Heather R. Manring, Saikh Jaharul Haque, and Arnab Chakravarti. 2024. "“Oh, Dear We Are in Tribble”: An Overview of the Oncogenic Functions of Tribbles 1" Cancers 16, no. 10: 1889. https://doi.org/10.3390/cancers16101889

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