Next Article in Journal
Reducing Hospital Readmissions in Chronic Obstructive Pulmonary Disease Patients: Current Treatments and Preventive Strategies
Previous Article in Journal
Etiological Aspects of Infectious Endocarditis in a Tertiary Hospital in Northeastern Romania
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mitochondrial Ribosomal Proteins and Cancer

1
Department of Pathology, The Affiliated Taizhou People’s Hospital of Nanjing Medical University, Taizhou 225300, China
2
Translational Medicine Center, The Affiliated Taizhou People’s Hospital of Nanjing Medical University, Taizhou 225300, China
*
Author to whom correspondence should be addressed.
Medicina 2025, 61(1), 96; https://doi.org/10.3390/medicina61010096
Submission received: 27 November 2024 / Revised: 19 December 2024 / Accepted: 7 January 2025 / Published: 9 January 2025
(This article belongs to the Section Oncology)

Abstract

:
Mitochondria play key roles in maintaining cell life and cell function, and their dysfunction can lead to cell damage. Mitochondrial ribosomal proteins (MRPs) are encoded by nuclear genes and are assembled within the mitochondria. MRPs are pivotal components of the mitochondrial ribosomes, which are responsible for translating 13 mitochondrial DNA-encoded proteins essential for the mitochondrial respiratory chain. Recent studies have underscored the importance of MRPs in cancer biology, revealing their altered expression patterns in various types of cancer and their potential as both prognostic biomarkers and therapeutic targets. Herein, we review the current knowledge regarding the multiple functions of MRPs in maintaining the structure of the mitochondrial ribosome and apoptosis, their implications for cancer susceptibility and progression, and the innovative strategies being developed to target MRPs and mitoribosome biogenesis in cancer therapy. This comprehensive overview aims to provide insights into the role of MRPs in cancer biology and highlight promising strategies for future precision oncology.

1. Introduction

The mitochondrion, a semi-autonomous organelle, possesses unique machinery for transferring information from its DNA (mtDNA) to proteins. This intricate process enables the mitochondrion to fulfill vital functions such as replication, transcription, and translation, thereby maintaining its genetic identity and supporting its pivotal role in cellular metabolism [1]. Human mitochondria contain approximately 1500 proteins, of which only 13 are encoded by mtDNA [2]. The vast majority of these proteins are encoded by nuclear DNA and are, subsequently, imported into the mitochondria. Therefore, mitochondrial genome evolution is characterized by a tight co-evolution with the nuclear genome, driven by the necessity for coordinated expression of mitochondrial and nuclear-encoded proteins essential for mitochondrial function [3,4,5]. This process is influenced by high mutation rates in mtDNA due to factors such as lack of protective histones and limited DNA repair mechanisms, leading to significant implications for aging and disease [3,6]. The dynamic interplay between nuclear and mitochondrial genomes shapes the somatic evolution of mitochondrial DNA throughout an organism’s lifetime [3].
Mitoribosomes play a pivotal role in the life activities of cells by orchestrating the synthesis and translation of 13 proteins vital for maintaining mitochondrial function and energy metabolism. This process is fundamental to the normal survival and metabolic activity of cells, ensuring their ongoing health and vitality. Human mitoribosomes are composed of a small 28S subunit and a large 39S subunit, together forming a functional 55S ribosome [7]. These mitoribosomes consist of RNA molecules and mitochondrial ribosomal proteins (MRPs), which are encoded by the mitochondrial and nuclear genomes, respectively [8]. The MRPs, encoded by nuclear genes and synthesized by the cytoplasm 80S ribosomes, contain 30 small mitoribosomal proteins (MRPSs) and 52 large mitoribosomal proteins (MRPLs) [9]. The review briefly outlines the structure of mitoribosomes and the function of MRPs, as well as their involvement in cancer.

2. MRPs and Mitoribosome Structure

Unlike cytoplasmic ribosomes encapsulated by a membrane, mitoribosomes lack such an encapsulation and have an irregular granular appearance [10]. In comparison to bacterial ribosomes and eukaryotic cytoplasmic ribosomes, both of which typically consist of 33% protein and 67% RNA, the ratio of MRP to RNA in mitoribosomes is notably reversed, falling approximately within the range of 25% to 30% [11,12]. Mammalian mitoribosomes have a low sedimentation coefficient of 55S, consisting of 28S small subunits (mtSSU) and 39S large subunits (mtLSU). The mtSSU is composed of 12S mitochondrial rRNA and 32 MRPs and provides a platform for mRNA binding and decoding. It is pivotal to translate leaderless mRNA, meticulously selecting translation initiation sites and exhibiting conformational adaptability during the translation process. In contrast, the mtLSU is composed of 16S mitochondrial rRNA and 50 MRPs [2]. The mtLSU plays a pivotal role in the catalytic peptidyl transferase reaction, which is essential for protein synthesis within the mitochondria. Its attributes include intricate interactions with mitochondria tRNA molecules to enable the accurate assembly of amino acids into proteins, the adaptation of the exit tunnel to ensure a smooth passage for hydrophobic nascent peptides emerging from the ribosomes, and the extensive remodeling of the central projection to facilitate the recruitment of specific tRNA molecules such as mitochondrial valine transfer RNA.
Unlike the cytoplasmic ribosomes, mitoribosomes are, instead, anchored to the inner mitochondrial membrane through the MRPL45 of their large subunit [2]. Mitoribosome biogenesis takes place in specific compartments within the mitochondrial matrix near the mtDNA nucleoid termed RNA granule [13]. This compartment contains RNA-processing enzymes, MRPs, and other essential proteins that are pivotal for mitoribosome biogenesis (Figure 1). Given that mtDNA encodes only 13 proteins, mitochondria rely heavily on the nucleus and various cellular compartments as the primary source for the vast majority of their proteins and lipids [2]. This intricate interplay underscores the interdependence between these two genetic systems in sustaining the vital functions of mitochondria. These intricately orchestrated nuclear-encoded subunit proteins are synthesized in the cytosol and imported into the mitochondria for their subsequent assembly. The transmembrane transport process of these proteins is a highly orchestrated event, mediated by chaperones and the sophisticated transmembrane transport complex, including components such as the translocase of outer mitochondrial membrane (TOM) and the translocase of the inner membrane (TIM) [14]. In addition, these 13 proteins are indispensable for the proper functioning of the oxidative phosphorylation (OXPHOS) system [15,16]. MRPs play an indirect yet crucial role in energy metabolism by facilitating the synthesis of 13 proteins that are integral components of the electron transport chain (ETC) and the ATP synthase complex, both of which are central to the OXPHOS process [17].
Cryo-electron microscopy (cryo-EM) has indeed revolutionized structural biology by enabling the determination of macromolecular structures at a near-atomic resolution [18]. This level of detail allows researchers to visualize not only the overall architecture, but also the fine features of complex biological assemblies such as mitoribosomes [19]. By capturing high-resolution snapshots of mitoribosomes in various functional states, cryo-EM provides indispensable insights into the mechanisms underlying mitochondrial translation and offers a powerful platform for investigating the molecular basis of mitoribosome-related diseases. The mtSSU is divided into three regions, namely head, platform, and foot. The head region of the mtSSU is characterized by the presence of death-associated protein 3 (DAP3), which interacts with other ribosomal proteins and rRNA to form a stable structure [20,21,22]. It forms two bridges with MRPL46 and MRPL48 by interacting with the central protuberance (CP) of the mtLSU. The foot region of the mtSSU is characterized by the presence of the pentapeptide repetition (PPR) domain protein of MRPS27, which is thought to play a role in the recognition and binding of specific RNA molecules. The interaction between MRPS27 and the N-terminal extension of MRPL19 forms a bridge that physically links the mtSSU and mtLSU. These interactions contribute to maintain the three-dimensional conformation of the mitochondrion, which is essential for its proper functioning. The structure of the mtLSU is more complex and larger than that of the mtSSU, and it also contains both rRNA molecules and numerous ribosomal proteins. The mtLSU accommodates tRNA-binding sites, the L1 stalk, the P stalk essential for recruiting translation factors, the peptidyl transferase center (PTC) that serves as the catalyst for the formation of peptide bonds between amino acids, the CP that facilitates seamless communication among various functional sites, and the exit tunnel through which the synthesized protein is egressed. The assembly of mitoribosomes is a highly coordinated and complex process that involves an extensive network of assembly factors and GTPases. To date, there have been few studies focused on the assembly pathway of the mammalian mtSSU [23,24]. The assembly of the mtSSU encompasses factors including ERAL1, NOA1/MTG3, MCAT, METTL17, TFB1M, RBFA, METTL15, and MTIF3 [19]. In contrast, the assembly of the mtLSU involves a different set of factors, including DDX28, GTPBP10, MRM3, the MALSU1-LOR8F8-mt-ACP module, NSUN4, MTERF4, MRM2, GTPBP7/MTG1, GTPBP5/MTG2, MtEF-TW/TUFM, and GTPBP6. Although the entrance to the human mitochondrion is wider than that of bacteria due to the absence of uS4 and the deletion of the C-terminal domain of uS3, the diameter of the mRNA channel is still smaller than that of an RNA duplex, thereby allowing only single-stranded mRNA to enter [21].

3. MRPs and Apoptosis

In addition to their role in energy production, mitochondria play a crucial role in apoptosis or programmed cell death. Alterations in mitochondrial function and defects in apoptosis are well-recognized hallmarks of cancer cells [25]. Dysregulation of MRPs can lead to mitochondrial dysfunction, which disrupts cellular energy metabolism and the balance of pro- and anti-apoptotic factors, thereby contributing to the survival and proliferation of cancer cells [17,26]. Specifically, MRPs such as DAP3, MRPL41, and MRPS30 are involved in the regulation of apoptosis.
DAP3 is a multifunctional protein that not only contributes to the assembly and function of the mtSSU, but also acts as a pro-apoptotic factor by interacting with other proteins to promote cell death [27]. DAP3 actively participates in the extrinsic apoptotic pathway, triggered by the tumor necrosis factor (TNF), the Fas ligand, and the TNF-related apoptosis-inducing ligand (TRAIL) [28,29]. As shown in Figure 2, DAP3 directly binds to the death domains (DD) of death receptors (DRs), including TNF receptor type 1 and TRAIL receptors DR4 and DR5 [27,30]. This binding enables DAP3 to form a complex with caspase-8 through Fas-associated death domain proteins, thereby initiating the apoptotic process. DAP3 also interacts with DAP3 binding cell death enhancer 1 (DELE1) to coordinate apoptotic events [30]. Notably, DELE1 knockdown enhances cell viability and inhibits DR-mediated apoptosis in response to the stimulation of TNF, anti-Fas, or TRAIL. Furthermore, liver kinase B1 (LKB1) has been shown to facilitate TRAIL-mediated apoptosis through the interaction of LKB1 interacting protein 1 with DAP3 [31]. This suggests a role for LKB1 in enhancing the sensitivity of cells to TRAIL-induced apoptosis through this specific protein–protein interaction network. Moreover, DAP3 is a key player in anoikis, the detachment-induced apoptosis of epithelial cells, through its interaction with the FAK-caspase-8 axis [32]. The regulation of the pro-apoptotic function of DAP3 is complex, modulated by various factors including AKT/PKB-mediated phosphorylation, transcriptional regulation of HML-10 transcripts, and post-transcriptional regulation by miR-365-1 [33,34,35]. These various regulatory mechanisms highlight the intricate balance that governs the function of DAP3 in cellular processes, particularly in the context of apoptosis.
MRPL41, also known as Bcl-2 interacting mitochondrial ribosomal protein (BMRP), is a component of the mtLSU. MRPL41 has the ability to trigger apoptosis through a p53-dependent pathway, with Bcl-2 and caspases serving as key mediators [36]. The 13–17 amino acid region of MRPL41 is necessary for binding to Bcl-2, with aspartic acid residue 16 being essential for this interaction. In addition, MRPL41 enhances the stability of p53 and facilitates its translocation to the mitochondria, thereby inducing p53-dependent apoptosis. The apoptosis-inducing ability of MRPL41 can be inhibited by Bcl-2 or by the caspase inhibitor. A recent study by Guo et al. [37] reported that inhibition of MRPL41 enhanced cell viability and increased Bcl-2 expression, in addition to suppressing apoptosis and downregulating p53. These multifaceted effects of MRPL41 underscore its significant role in regulating cell survival and cell death pathways. MRPS30, also known as programmed cell death protein 9 (PCDP9), is a component of the mtSSU. Although MRPS30 has also been implicated in the regulation of apoptosis through the activation of c-Jun N-terminal kinase 1 (JNK1) in murine fibroblasts [38], the exact mechanisms by which MRPS30 contributes to apoptosis are not as well characterized as those of MRPL41.

4. MRPs Associated with Cancers

Cancer is a complex disease characterized by its diverse manifestations and underlying causes, resulting from perturbations in multiple cellular processes. These alterations include the disruption of growth signals, the evasion of apoptosis, unlimited replication potential, angiogenesis, invasion, and metastasis [25]. In 1926, Otto Warburg first described the differences in the glycolytic production of ATP under aerobic conditions between tumor cells and normal cells, a phenomenon now known as the Warburg effect [39]. Mitochondrial ribosomes play a crucial role in mitochondrial function. In recent years, numerous studies have uncovered potential mechanisms by which MRPs are involved in tumorigenesis. MRPs are closely linked to the occurrence and development of cancer. Therefore, the characterization and understanding of dysregulated MRPs in cancer may provide a powerful tool for cancer diagnosis, prognosis, and treatment outcomes.

4.1. MRPs and Lung Cancer

Lung cancer is the most common type of cancer and the primary cause of cancer-related deaths globally [40]. Given that metabolic reprogramming is a key feature of lung cancer, the dysregulation of MRPs can lead to altered energy metabolism, thereby supporting the rapid growth and survival of cancer cells [17,41]. Bennett et al. [42] found that silencing mitochondrial ribosomal proteins and respiratory chain genes in primary and metastatic lung cancers lead to significant changes in mitochondrial respiration, with primary tumors showing a greater dependence on these genes for growth compared to metastatic tumors. This suggests that primary tumors are more sensitive to disruptions in mitochondrial function, whereas metastatic tumors may have developed alternative metabolic pathways. The energy metabolic phenotypes of lung cancer are not singular but rather represent a spectrum of metabolic states that evolve as the disease progresses [42,43]. In addition, the gene expression profiling interactive analysis (GEPIA) database reveals differential expression of MRPs in lung cancers [44]. Specifically, eight MRPs are significantly upregulated in lung adenocarcinomas (LUADs), while, in lung squamous cell carcinomas (LUSCs), 22 MRPs are upregulated and 1 MRP is downregulated (Figure 3). Notably, seven MRPs (MRPL12, MRPL3, MRPL15, MRPL24, MRPS23, MRPS33, and MRPS35) have been found to be all upregulated in both LUADs and LUSCs. This indicates that these shared MRPs may play a fundamental role in mitochondrial function across different subtypes of lung cancer, potentially indicating common metabolic adaptations or vulnerabilities that could be targeted therapeutically.
MRPL12 is a key player in mitoribosome biogenesis and transcription, interacting with mitochondrial RNA polymerase to activate mtDNA transcription and regulate gene expression [45]. Therefore, MRPL12 is essential for maintaining mitochondrial homeostasis, particularly in the regulation of mitochondrial respiration [46]. MRPL12 knockdown induces structural damage in mitochondria, characterized by swelling, irregular arrangement, less recognizable cristae, and reduced volume and abundance. MRPL12 is upregulated in many types of cancer, including lung cancer, hepatocellular carcinoma (HCC), and breast cancer [46,47]. MRPL12 enhances cell proliferation, migration, and invasion of lung cancer cells by upregulating mitochondrial oxidative phosphorylation, and its expression levels are inversely associated with the infiltration levels of multiple immune cells. Similar findings have also been observed in HCC, where MRPL12 promotes tumor growth and metastasis through the regulation of mitochondrial metabolism [45]. High expression levels of both MRPL12 mRNA and protein are associated with a poor prognosis in lung cancer as well as HCC [45,47]. The MRPL12 Y60 site is a specific phosphorylation modification site that regulates its oncogenic functions [46]. Furthermore, Liu et al. [48] observed an oncogenic role of MRPL12 that promoted cell proliferation and migration of breast cancer cells. High MRPL12 expression was weakly but significantly associated with a poor prognosis in breast cancer. These results indicate that MRPL12 may be a promising prognostic biomarker and therapeutic target for lung cancer, breast cancer, and HCC.
MRPL13 has been involved in various biological processes, including participation in protein synthesis within the mitochondrion, the adaptation of the translation system to the specific requirements of the organelle, and the structural and functional integrity of the mitochondrion [49,50]. Furthermore, many studies have demonstrated that MRPL13 has a significant role in cancer. MRPL13 is dysregulated in many types of cancer, including breast cancer, colorectal cancer (CRC), gastric cancer, and lung cancer [51,52]. High MRPL13 expression is associated with a worse prognosis in several types of cancer, including lung cancer, breast cancer, and gastric cancer [51,52,53,54]. There is an inverse association between MRPL13 expression and the abundance of tumor-infiltrating lymphocytes (TILs), such as B cells, CD8+ T cells, and macrophages [51]. MRPL13 expression influences the sensitivity to anticancer drugs, including tramadol, CI-1040, navitoclax, and nutlin-3A. Zhong et al. [51] observed that MRPL13 knockdown promoted the apoptosis and inhibited division, metastasis, and invasion in LUADs. Cai et al. [53] reported that MRPL13 promoted breast cancer invasion, metastasis, and epithelial–mesenchymal transition (EMT) through the PI3K/AKT/mTOR signaling pathway, highlighting its involvement in cancer progression. High MRPL13 expression was associated with adverse clinicopathological variables and unfavorable clinical outcomes in breast cancer patients. A recent study by Dai et al. [54] also demonstrated that MRPL13 facilitated the invasion of breast cancer cells. The expression levels of MRPL13 in metastatic breast cancer tissues were obviously higher than those in primary breast cancer tissues. Wang et al. [52] observed that MRPL13 promoted cell proliferation and G1/S phase transition in gastric cancer cells and enhanced xenograft growth by suppressing the p53 signaling pathway. Lee et al. [55] reported that exogenous lactate inhibited mitochondrial OXPHOS by downregulating MRPs, particularly MRPL13, thereby facilitating invasion of HCC cells through the induction of claudin-1. Taken together, MRPL13 is a diagnostic and prognostic biomarker for several types of cancer, including lung cancer, breast cancer, and gastric cancer [51,53,54].
MRPL9 is a component of the mtLSU and is upregulated in many types of cancer, including lung cancer, HCC, and breast cancer [56]. Li et at. [57] observed that MRPL9 promoted cell proliferation, migration, and invasion of lung cancer cells by regulating c-MYC and its overexpression was associated with a poor prognosis in lung cancer patients, indicating that MRPL9 may be a potential therapeutic target for lung cancer. Upregulation of MRPL15 is observed in many types of cancer, including lung cancer, CRC, and gastric cancer [58]. The expression levels of MRPL15 are inversely correlated with the infiltration levels of multiple immune cells. High expression of both MRPL15 mRNA and protein is associated with a poor prognosis in lung cancer patients. Similarly to MRPL15, upregulation of MRPL19 is observed in many types of cancer, including lung cancer, CRC, and gastric cancer [59]. Wei et al. [59] found that MRPL19 knockdown induced apoptosis and inhibited the migration and invasion of lung cancer cells. The expression levels of MRPL19 were inversely correlated with the infiltration levels of B cells, CD4+ T cells, and dendritic cells. High expression of both MRPL19 mRNA and protein were associated with a poor prognosis in lung cancer patients. MRPL51 is crucial for mitochondrial genome stability, with a key role in redox homeostasis [60]. Lack of MRPL51 leads to loss of mitochondrial genome and impaired oxidative phosphorylation, reducing superoxide production. A recent study by Zhang et al. [61] reported that MRPL51 was transcriptionally activated by FOXM1 and its mRNA, and that protein levels were upregulated in LUADs. At the single-cell level, the expression level of MRPL51 exhibited a positive correlation with various cellular processes in LUAD cells, including the cell cycle, DNA damage, DNA repair, EMT, invasion, and proliferation. Upregulation of MRPL51 contributed to the malignant behaviors of LUAD cells and was associated with a poor prognosis. Gu et al. [62] reported a risk model based on three differentially methylated genes, including MRPL51, SLK, and PRC1, which exhibited promise for guiding the management of LUAD treatment. Guttapadu et al. [63] found that MRPL52 was upregulated in LUAD and LUSC patients, and its overexpression was correlated with a poor prognosis in LUSC patients.

4.2. MRPs and Breast Cancer

Breast cancer is the most commonly diagnosed cancer and the fifth leading cause of cancer-related deaths among women in China [64]. Although MRPs are dysregulated in breast cancer patients, including the aforementioned MRPL9, MRPL12, and MRPL13, the specific mechanisms underlying the activity of these proteins in relation to this cancer have yet to be explored.
Lin et al. [65] analyzed the expression patterns and prognostic significance of MRPs in breast cancer patients using the GEPIA database. They identified 12 upregulated MRPs, including MRPL3, MRPL13, MRPL14, MRPL17, MRPL24, MRPL42, MRPL47, MRPS23, DAP3, MRPS30, MRPS34, and MRPS35. Notably, MRPL13, MRPL14, MRPL17, MRPL42, DAP3, and MRPS35 were upregulated across all four subtypes. A prognostic model including four MRPs (MRPL16, MRPL40, MRPS18C, and MRPS35) demonstrated a good performance in predicting the survival outcomes of breast cancer patients. Furthermore, we conducted an analysis of 1,113 breast cancer samples from The Cancer Genome Atlas (TCGA) to identify differentially expressed MRPs across various subtypes. Our findings revealed distinct expression patterns of MRPs that varied significantly between different breast cancer subtypes (Table 1). These expression patterns were inconsistent with the results reported by Lin et al. [65], something which may be attributable to differences in sample size. Five MRPs (MRPL12, MRPS12, MRPL13, MRPL14, and MRPS34) were upregulated across all five subtypes. There was subtype-specific upregulation of MRPs, including MRPL17, MRPS23, MRPL21, and MRPS28 in the luminal B subtype, and GADD45GIP1, MRPL2, MRPL4, MRPL9, MRPS21, MRPL36, MRPL37, and MRPL51 in the basal-like subtype. These subtype-specific patterns suggest potential oncogenic influences on cancer progression, highlighting the diverse roles of MRPs in different breast cancer subtypes.
MRPS23 is upregulated in various types of cancer, including breast cancer, HCC, pancreatic adenocarcinoma, and glioma [66]. The aberrant expression of MRPS23 is correlated with a poor prognosis in relation to many types of cancer, including breast cancer [66,67]. Although Gatza et al. [68] observed an association of MRPS23 amplification with poor prognosis in breast cancer patients, this finding was not validated by Klæstad et al. [69]. Two studies reported that MRPS23 knockdown in breast cancer cells suppressed proliferation, metastasis, and invasion while promoting apoptosis and increasing sensitivity to paclitaxel [70,71]. Furthermore, the phosphorylation of MRPS23 by mitotic kinases influenced cell proliferation, migration, and enhanced sensitivity to CDK1 inhibitors in breast cancer cells [67,72]. Liu et al. [67] found that methylation of arginine 21 and lysine 108 on MRPS23 coordinatively promoted breast cancer cell survival and malignancy by inhibiting OXPHOS. These findings suggest that the role of MRPS23 in breast cancer may be more complex and nuanced than initially thought. Furthermore, MRPS23 also contributes to the malignant phenotypes of other types of cancer, including glioma, HCC, and cervical cancer [66,73,74]. These findings indicate that MRPS23 may play a significant role in promoting the aggressive traits of these cancers. Further studies are required to fully elucidate its exact mechanisms in these diseases.
MRPS27 functions as an RNA-binding protein that interacts with 12S mitochondrial rRNA and participates in the translation of mRNAs [75]. A recent study by Zheng et al. [76] found that MRPS27 was upregulated in triple-negative breast cancer (TNBC). MRPS27 promoted cell proliferation, migration, invasiveness, and tumor growth in TNBC. Inhibition of MRPS27 triggered nucleolar stress and suppressed the stemness properties in TNBC. Upregulation of MRPS27 diminished the stemness-inhibitory effects of lovastatin in TNBC cells, indicating an important role of MRPS27 in maintaining the stemness properties. High MRPS27 expression was associated with advanced TNM stage and a worse prognosis in TNBC patients. Further studies are required to explore the mechanisms of MRPS27 and validate its prognostic significance in the context of breast cancer. MRPL52 is upregulated in breast cancer, particularly in metastatic cases [8]. Breast cancer patients with high MRPL52 expression have a poor prognosis. MRPL52 regulates hypoxia-induced apoptotic resistance and metastatic initiation of breast cancer cells by facilitating PINK1/Parkin-dependent mitophagy and transactivating the reactive oxygen species (ROS)-Notch1-Snail pathway, respectively.
Neoadjuvant chemotherapy (NAC) is the standard of care for locally advanced breast cancer and increases the rate of breast-conserving surgery. Vishnubalaji et al. [77] identified a gene signature including MRPL9 and MRPL37 associated with resistance to NAC in TNBC at the single-cell level. MRPL9 and MRPL37 were upregulated in TNBC resistant to NAC. Inhibiting MRPL37 lead to a reduction in TNBC colony formation and an enhancement in sensitivity to paclitaxel, but these effects were not observed when MRPL9 was inhibited. To date, there are limited reports on MRPL9 and MRPL37 in cancer, and our understanding of their roles in the disease is still limited.
Previous studies have demonstrated that the 5p12 variants rs10941679 and rs7716600 confer susceptibility to estrogen receptor (ER)-positive breast cancer [78,79,80]. The rs10941679 is located within an enhancer element that physically interacts with the promoter regions of FGF10 and MRPS30 [79]. Although there is no difference in transcription factor binding affinity between rs10941679 alleles, the G allele confers a higher expression of FGF10 and MRPS30. Furthermore, rs7716600 is the variant in the 5p12 region with the greatest effect on the expression level of MRPS30, and the risk A allele is associated with decreased methylation at the MRPS30 promoter site in ER-positive breast cancer cells [80]. It is interesting that MRPS30 is not expressed in normal breast luminal epithelial cells but is significantly upregulated in breast cancer patients, a phenomenon which contradicts its role in promoting apoptosis [38,81,82]. MRPS30 is dysregulated in many types of cancer, including breast cancer, CRC, lung cancer, and gastric cancer [17]. The mechanism of MRPS30 in the development and progression of ER-positive breast cancer is still not clearly understood.

4.3. MRPs and Digestive System Cancers

Gastric cancer is a prevalent malignant tumor of the gastrointestinal tract, ranking fifth for incidence and fourth for mortality, globally [40]. According to the GEPIA database, 16 MRPs are significantly upregulated in gastric cancer (Figure 4) [44]. These MRPs include MRPL3, MRPL13, MRPL14, MRPL15, MRPL17, MRPL19, MRPL30, MRPL42, MRPL47, MRPL48, MRPL57, MRPS30, MRPS10, MRPS17, and MRPS35. The significant upregulation of these MRPs indicates their potential involvement in altered mitochondrial function and metabolism in gastric cancer cells. MRPS17 is upregulated in many types of cancer, including gastric cancer [83,84,85]. Upregulated MRPS17, along with the elevated expression of other MRPs, enhances OXPHOS in cancer cells, thereby fueling cancer metabolism [86]. Zhou et al. [83] reported that MRPS17 facilitated the migration and invasion of gastric cancer cells by regulating the PI3K/AKT signaling pathway. High MRPS17 expression was associated with a worse prognosis in gastric cancer patients. Li et al. [87] developed a seven-gene signature including MRPS17 for predicting the prognosis of patients with gastric cancer. The expression level of MRPS17 was inversely correlated with the abundance of TILs, including Th1 cell, Tem-CD8 cell, NKT cell, and NK cell. A recent study by Hu et al. [88] identified a three-gene signature including MRPS17, GUF1, and PDK4 for predicting the prognosis of gastric cancer patients. These findings imply that MRPS17 may serve as a prognostic biomarker for gastric cancer. MRPL35 is involved in the formation of the CP of the mtLSU, and its interaction with other MRPs such as MRPL38 is essential for optimizing the mitotranslational process to ensure the synthesis of competent proteins for the assembly into functional OXPHOS enzymes, particularly cytochrome c oxidase [89,90]. MRPL35 is upregulated in several types of cancer, including lung cancer, CRC, gastric cancer, and HCC, and its overexpression is associated with a worse prognosis in these cancer types [91,92,93,94,95]. High MRPL35 expression is correlated with an advanced stage of gastric cancer [92]. MRPL35 knockdown induces apoptosis in gastric cancer cells, inhibits their proliferation and colony formation, and suppresses tumor growth by regulating BCL-XL, PICK1, and AGR2. 18β-glycyrrhetinic acid (18β-GRA), extracted from the natural medicine licorice, has demonstrated reliable anticancer effects on human cancer cells [96]. 18-GRA inhibits cell proliferation and tumor growth and promotes apoptosis by downregulating MRPL35 in gastric cancer patients. MRPL35 knockdown inhibits cell proliferation and tumor growth in CRC by triggering the accumulation of ROS, which subsequently leads to DNA damage, thereby suppressing the disease progression in CRC [91]. There is a significantly lower level of the promoter methylation of MRPL35 in HCC, a phenomenon which is consistent with its upregulation in HCC [95]. MRPL35 promotes migration and invasion of HCC cells, confers resistance to anticancer drugs, and likely exerts negative effects on the immune system. Wang et al. [97] reported that MRPL35 knockdown inhibited cell proliferation and induced apoptosis in esophageal carcinoma (ESCA) TE-1 cells. In addition, MRPL35 facilitates cell proliferation, invasion, glutamine metabolism, and tumor growth by regulating SLC7A5 in lung cancer patients, whereas it is deubiquitinated and stabilized by ubiquitin-specific protease 39 [93]. MRPL35 may play important roles in cancer progression and confer multidrug resistance, particularly in relation to gastric cancer and lung cancer. Sotgia et al. [98] identified 12 MRPLs and eight MRPSs that showed a significant prognostic value and developed an eight-gene signature including MRPL28 and MRPL52 for predicting the prognosis in gastric cancer patients. A more recent study by Wang et al. [99] reported that a nine-gene signature including MRPL4 served as an independent indicator for the prognosis in gastric cancer patients and was associated with immune escape mechanisms as well as the response to PD-L1 immunotherapy. Further research into the specific mechanisms by which these MRPs contribute to cancer progression could provide valuable insights and lead to the identification of novel therapeutic strategies targeting mitochondrial pathways.
HCC is the fourth most common cancer and the second leading cause of cancer-related deaths in China [64]. Many MRBs are associated with the progression of HCC, including the aforementioned MRPL9, MRPL12, MRPL13, MRPS23, and MRPL35. A recent study by Zhao et al. [100] identified 14 differentially expressed MRPs (MRPS17, MRPS21, MRPS23, MRPL9, DAP3, MRPL13, MRPL14, MRPL15, MRPL16, MRPL17, MRPL21, MRPL24, MRPL47, and MRPL55) in HCC from TCGA, all of which exhibited a good diagnostic performance. Furthermore, the study also revealed that 39 MRPs were associated with a better or worse prognosis in HCC. Previous studies have demonstrated that MRPL21 may potentially play a role in several critical cellular processes, including the activation of the cell cycle, the enhancement of apoptotic pathways, and the triggering of DNA damage responses [100]. These multifaceted functions highlight the importance of MRPL21 in maintaining cellular homeostasis and responding to stress. MRPL21 is dysregulated in some types of cancer, including HCC and ESCA [100,101,102]. The aberrant expression of MRPL21 is associated with a poor prognosis of HCC and with the drug resistance of acute myeloid leukemia (AML) and LUSC [101,103,104]. Upregulation of MRPL21 promotes cell proliferation and confers resistance to the apoptosis of HCC cells, which may be responsive because of the increased incidence of TP53 mutations [101]. Nutlin-3, a P53 inhibitor, effectively potentiates the pro-apoptotic effects of MRPL21 knockdown in HCC cells. These findings highlight the importance of MRPL21 in the progression of HCC and its potential as a prognostic biomarker and therapeutic target. MRPL48 is upregulated in many types of cancer, including HCC, breast cancer, and gastric cancer, a phenomenon which may be due to the hypomethylation of its promoter [105]. Dysregulation of MRPL48 has been observed in HBV-, non-alcoholic steatohepatitis-related HCC, but not in HCV-related HCC. MRPL48 promotes the cell proliferation, migration, and invasion of HCC cells and influences immune infiltration levels in HCC. High MRPL48 expression is associated with a more aggressive behavior of HCC and a worse prognosis. Hu et al. [106] reported that MRPL48 knockdown inhibited cell proliferation and enhanced cetuximab sensitivity in CRC. However, there is limited research and understanding regarding the role of MRPL48 in cancer, and its role in cancer is not well understood. Further studies are required to elucidate the mechanisms by which MRPL48 contributes to cancer development and progression.
CRC is the second most common cancer and the fourth leading cause of cancer-related deaths in China [64]. Previous studies have revealed that some MRPs are dysregulation in CRC, including the aforementioned MRPL9, MRPL19, MRPL13, MRPL15, MRPL35, and MRPL48. Building on this foundation, our analysis using the GEPIA database identified 19 MRPs as being significantly upregulated in colon adenocarcinomas (COADs), including MRPL3, MRPL12, MRPL13, MRPL14, MRPL15, MRPL17, MRPL19, MRPL30, MRPL37, MRPL42, MRPL48, MRPL57, MRPS12, MRPS16, MRPS23, MRPS28, MRPS33, MRPS34, and MRPS35 (Figure 4) [44]. Similarly, 18 MRPs are significantly upregulated in rectum adenocarcinomas (READs), including MRPL3, MRPL13, MRPL14, MRPL15, MRPL17, MRPL19, MRPL30, MRPL37, MRPL42, MRPL48, MRPL57, MRPS12, MRPS17, MRPS23, MRPS28, MRPS33, MRPS34, and MRPS35. The alternative splicing of MRPL33 pre-mRNA is regulated by hnRNPK, leading to long (L) and short (S) isoforms [107]. The MRPL33-L abundance in cancer tissues is increased compared to that in normal tissues, including CRC and gastric cancer [107,108]. MRPL33-L exhibits an oncogenic function that facilitates cell proliferation and inhibits apoptosis in CRC, whereas MRPL33-S does not affect these phenotypes [107]. Furthermore, Li et al. [108] found that MRPL33-L and MRPL33-S play distinct roles in the response of gastric cancer cells to chemotherapy. Specifically, MRPL33-L promoted resistance to epirubicin, while MRPL33-S enhanced sensitivity to epirubicin through the PI3K/AKT signaling pathway. MRPL43 is weakly but significantly upregulated in CRC [109]. MRPL43 knockdown inhibits cell proliferation, invasion, and migration, and promotes apoptosis in CRC. Abdul Aziz et al. [110] developed a 19-gene signature including MRPL52 that was an independent predictor of survival of CRC patients. MRPL41 is dysregulated in many types of cancer, including CRC, ESCA, and lung cancer [17]. Xin et al. [111] established a nine-gene signature including MRPL41 for the prediction of postoperative recurrence in stage II/III CRC.
To date, there are very few studies specifically investigating the role of MRPs in pancreatic adenocarcinomas. Chen et al. [112] developed a four-gene signature including DNAH10, HSBP1L1, KIAA0513, and MRPL3 for predicting the prognosis of pancreatic adenocarcinoma patients from The Cancer Genome Atlas (TCGA) dataset. Patients with a high-risk score based on the signature experienced a significantly shorter overall survival time. To validate these findings, the prognostic model was tested using an independent database, confirming its robustness and predictive accuracy. The inclusion of MRPL3 in the prognostic model underscores the potential importance of MRPs in pancreatic adenocarcinomas. Interestingly, the GEPIA database has also revealed that 63 MRPs, including MRPL3, are significantly upregulated in pancreatic adenocarcinomas (Table 2) [44]. This extensive list of upregulated MRPs suggests that these proteins might play critical roles in the disease’s progression and highlights the need for further investigation into their specific contributions. Such research could lead to the identification of novel biomarkers or therapeutic targets in pancreatic adenocarcinomas.
MRPL37 is dysregulated in some types of cancer, including ESCA, CRC, and breast cancer [113]. Liu et al. [113] developed a six-gene signature including MRPL37 that was an independent indicator for ESCA. The signature was also associated with resistance to chemoradiotherapy by modulating the immune microenvironment and impacting angiogenesis. Notably, only MRPL3 and MRPL47 are upregulated in ESCA according to the GEPIA database (Figure 4) [44]. Tian et al. [114] found that MRPS18A was upregulated in cholangiocarcinomas and that its overexpression was associated with a worse prognosis. A recent study by Kim et al. [115] report that upregulated MRPS18A was associated with an improved prognosis in patients with mismatch repair-proficient advanced refractory CRC who were treated with regorafenib and nivolumab. Moreover, increased expression of MRPS18A has also been observed in HCC and breast cancer patients [116,117]. According to the GEPIA database, MRPS18A exhibits dysregulation in pancreatic adenocarcinomas, ovarian serous cystadenocarcinomas, and AML (Table 2) [44]. Zhuang et al. [118] reported that MRPL27 was upregulated in cholangiocarcinomas and that it served as an independent prognostic biomarker. Huang et al. [119] identified a four-gene signature including SDHAF2, MRPS34, MRPL11, and COX8A that served as an independent indicator for patients with cholangiocarcinomas treated with adjuvant transarterial chemoembolization. These four key genes are critical for maintaining the stemness of cholangiocarcinoma stem cells and may confer chemoresistance to cholangiocarcinoma cells.
Abnormal expression of different MRPs in the same cancer type or of specific MRPs across different cancers can significantly impact cancer behavior and progression. However, it remains unclear whether there are differences in the mechanisms by which these MRPs influence cancer behavior.

4.4. MRPs and Other Cancers

In addition to the aforementioned breast cancer, lung cancer, and digestive tract cancers, there are other malignancies associated with MRPs. MRPL9 is upregulated and promotes cell proliferation, metastasis, and growth by regulating the MAPK/ERK signaling pathway in papillary thyroid cancers (PTCs) [120]. The interaction between γ-glutamylcyclotransferase and MRPL9 enhances the oncogenic function of MRPL9. PTC patients with elevated MRPL9 expression tend to have a worse prognosis. MRPL15 is upregulated in ovarian cancer, a phenomenon which may be partially due to the amplification and hypomethylation of MRPL15 [121]. MRPL15 influences the abundance of infiltrating immune cells, such as CD8+ T cells and activated dendritic cells. High MRPL15 expression is associated with an advanced stage and a poor prognosis in ovarian cancer patients. MRPL51 and MRPL52 are upregulated in ovarian cancer patients, with the low expression of both MRPL51 and MRPL52 associated with a worse prognosis [122]. Moreover, MRPL51 and MRPL52 influence the abundance of infiltrating immune cells, such as γδ T and memory B cells. The transcription of MRPL52 is regulated by LINC00152 and is found to be upregulated in oral squamous cell carcinomas (OSCCs) [123]. MRPL52 promotes OSCC cells proliferation, and its overexpression is associated with a worse prognosis [123]. Gou et al. [124] found that MRPL33 was upregulated in HPV-related OSCCs, and its expression level was further increased when treated with the bromodomain inhibitor, JQ1. Wei et al. [125] identified 19 differentially expressed MRPSs (|log2 (FC)| > 1, p < 0.05.) out of 25 MRPSs analyzed in clear cell renal cell carcinomas (ccRCCs). Eight MRPSs (MRPS14, MRPS15, MRPS26, MRPS28, MRPS30, MRPS31, MRPS35, MRPS36) were associated with overall survival. They subsequently developed a six-MRPS signature (MRPS14, MRPS15, MRPS26, MRPS28, MRPS30, and MRPS36) for predicting the prognosis of ccRCC patients. Furthermore, two potential therapeutic targets and eight agents were identified that might enhance the therapeutic effects on patients with high MRPS-related scores. A recent study by Yang et al. [126] found that MRPS5 was significantly upregulated in ccRCC patients. However, high MRPS5 expression was associated with a better prognosis in ccRCC patients, contrasting sharply with the poor prognosis observed in HCC patients with high MRPS5 expression. This discrepancy suggests the context-dependent roles of MRPS5 across different cancer types and underscores the need for further research to understand its mechanisms in various oncological settings. In addition, further studies are required to validate the prognosis significance of MRPS5 in ccRCC patients.
Hu et al. [127] established a six-gene signature including MRPL47 for predicting the prognosis of patients with squamous cell carcinomas of the head and neck. MRPL47 rs10513762 is a potential risk factor for vincristine-induced peripheral neuropathy in childhood AML [128]. A recent study by Liu et al. [129] reported that the elevated expression of both MRPL4 and MRPS12 was significantly associated with unfavorable clinical outcomes in AML patients. Fu et al. [130] found that MRPL4 was upregulated in high-risk prostate cancer patients, and patients with elevated MRPL4 expression had a worse prognosis. MRPS23 exhibits oncogenic functions in glioma patients, promoting the cell proliferation and metastasis of glioma cells, and serves as an independent prognostic indicator [66]. Upregulation of MRPS17 in glioblastomas confers resistance to chemotherapy including temozolomide and nitrosoureas [131]. Liu et al. [132] found that MRPS7 was upregulated in osteosarcoma patients and that it might be associated with metastasis. Bell et al. [133] reported that neuroblastoma patients with elevated MRPL11 expression had a worse prognosis. Bae et al. [134] found that MRPS18B rs148828689 was significantly associated with the risk of neuroblastoma in Korean children. According to the GEPIA database, MRPS18B is dysregulated in lower-grade gliomas and AML (Table 2) [44]. Furthermore, a previous study has demonstrated that MRPS18B interacts with p53, indicating its potential role in p53-related pathways [135]. These findings indicate that MRPS18B might play important roles in cancer development and progression. Song et al. [136] found that MRPS18A gene fusions were more commonly detected in high-grade endometrial stromal sarcomas (MRPS18A-ZAN) and uterine leiomyosarcomas (MRPS18A-PDC-AS1), suggesting their potential as biomarkers for the differential diagnosis between these two types of uterine sarcomas. However, the evidence regarding the effect of other MRPs on different types of cancer is limited, warranting further investigation in the future.
Furthermore, the GEPIA database has revealed that these types of cancer exhibit distinct expression patterns of MRPs (Table 2) [44]. The vast majority of MRPs are upregulated across various cancer types, while a few MRPs show downregulation in relation to specific cancers. This widespread upregulation suggests an increased metabolic activity in cancer cells, often associated with enhanced energy production to support rapid cell proliferation. Interestingly, 59 abnormally expressed MRPs in AML patients are downregulated, a phenomenon which is a unique pattern compared to other cancers. This suggests that the dysregulation of MRPs in AML patients may involve a different set of mechanisms compared to other cancers. Additionally, MRPL1 has been observed to be only dysregulated in pancreatic adenocarcinoma patients, and other MRPs are dysregulated in relation to various types of cancer described in the article. This indicates that MRPs play critical roles in multiple cancer types and highlights the complexity of their involvement in oncogenesis.

5. MRPs in Cancer Therapy

Drug resistance significantly limits the effectiveness of cancer treatments and poses a major challenge to precision medicine. Numerous studies have demonstrated that the biosynthesis and translation of mitochondrial ribosomes are intricately related to tumor development, metabolism, and chemotherapy resistance [10,12,26,137,138,139]. In this context, exploring the roles of MRPs in cancer patients offers promising potential for enhancing early diagnosis, improving prognosis, and developing innovative therapeutic strategies.
Accumulating evidence indicates that MRPs can impact multiple types of cancer, with some MRPs showing abnormal expression in relation to some specific types of cancer [17,100,125]. Therefore, MRPs have the potential to serve as therapeutic targets and diagnostic and prognostic biomarkers for various cancers. As mentioned above, certain MRPs, including MRPL33, MRPL37, MRPL48, and MRPS23, have been observed to confer resistance to anticancer drugs at the cellular level [72,77,95,106]. The expression levels of MRPL21 and MRPL35 are related to resistance to anticancer drugs [95,103,104]. MRPS17 and several gene signatures containing MRPs such as MRPL37, MRPS34, and MRPL11 may be independent indicators for cancer patients treated with chemotherapy or chemoradiotherapy [113,119,131]. In addition, JG-98, an allosteric inhibitor of the mitochondrial heat shock protein 70, re-sensitizes castration-resistant prostate cancer cells to androgen-deprivation therapy in vitro and in vivo by reducing MRP levels and OXPHOS activity [140]. SAMMSON modulates the RNA–protein complex formation in both cytosol and mitochondria, thereby promoting a balanced enhancement of rRNA processing and protein synthesis in these cellular compartments, a phenomenon which facilitates cancer cell growth [141]. Inhibition of SAMMSON using an antisense oligonucleotide induces apoptosis and suppresses tumor growth in uveal melanoma patient-derived xenograft (PDX) models [142]. Disrupting the equilibrium through the inhibition of the mitoribosome biogenesis machinery is emerging as a promising approach in anticancer therapy.
The successful use of antibiotic agents targeting bacterial ribosomes highlights the potential for mitoribosome inhibitor use in cancer treatments. Tetracycline analogs, such as tigecycline, doxycycline, and COL-3, inhibit mitochondrial protein synthesis by binding to the mtSSU, leading to reduced OXPHOS activity and to the suppression of cancer cell proliferation [143,144,145]. Tigecycline has demonstrated anticancer efficacy across various cancer types, including AML, HCC, lymphoma, gastric, ovarian, and breast cancers [139,146]. Notably, it has been observed to synergize with chemotherapy and other agents, such as pyrvinium pamoate, enhancing its therapeutic potential [143,147,148]. In vivo studies have shown that treatment with tigecycline or doxycycline can effectively inhibit the metastasis of OSCC [149]. These findings suggest potential therapeutic benefits of these agents in preventing the spread of OSCC and improving patient outcomes. However, the translation from preclinical experiments into practical applications presents substantial challenges [139]. For example, a Phase 1 clinical trial assessing tigecycline as a single agent in AML patients demonstrated no significant pharmacodynamic changes or clinical responses [150]. This suggests that achieving meaningful therapeutic effects with a single agent may be particularly difficult in AML patients, highlighting the need for alternative strategies such as combination therapies. Additionally, short-term pre-operative treatments with doxycycline have been shown to significantly reduce the number of cancer stem cells in breast cancer patients [151], further underscoring the promise and complexity of these therapeutic approaches.

6. Conclusions

Well-functioning mitochondria can support the proliferation and energy metabolism of normal cells. These MRPs are necessary for the composition of mitoribosomes, with or without a role in the assembly and translation of mitochondrial DNA. The distinct expression patterns of certain MRPs in relation to different cancers suggest their potential as diagnostic and prognostic biomarkers for these diseases. However, further studies are warranted to validate these findings. Targeting MRPs and mitoribosome biogenesis is emerging as a promising approach in anticancer therapy. Nevertheless, translating this potential into clinical practice will require extensive research to establish its efficacy and safety. To date, the specific mechanisms that trigger cancer development and promote its progression remain largely unknown. Investigating the functional roles of these MRPs across different types of cancer could uncover key mechanisms driving disease development and progression, as well as revealing the extra-ribosomal properties. This, in turn, could provide valuable insights and clues for the development of innovative therapeutic strategies.

Author Contributions

H.W., X.Z., H.Z., M.S., J.Y., and H.Y. reviewed the literature and contributed to the manuscript drafting; H.W. and H.Y. were responsible for the revision of the manuscript for important intellectual content. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Taizhou Social Development Project Foundation, Jiangsu, China (grant No. TS202309) and by the Project Foundation of Taizhou School of Clinical Medicine, Nanjing Medical University (grant No. TZKY20220204 and grant No. TZKY20220205).

Conflicts of Interest

The authors declare that they have no conflicts of interest.

References

  1. Gustafsson, C.M.; Falkenberg, M.; Larsson, N.G. Maintenance and Expression of Mammalian Mitochondrial DNA. Annu. Rev. Biochem. 2016, 85, 133–160. [Google Scholar] [CrossRef] [PubMed]
  2. Gopisetty, G.; Thangarajan, R. Mammalian mitochondrial ribosomal small subunit (MRPS) genes: A putative role in human disease. Gene 2016, 589, 27–35. [Google Scholar] [CrossRef] [PubMed]
  3. Serrano, I.M.; Hirose, M.; Valentine, C.C.; Roesner, S.; Schmidt, E.; Pratt, G.; Williams, L.; Salk, J.; Ibrahim, S.; Sudmant, P.H. Mitochondrial haplotype and mito-nuclear matching drive somatic mutation and selection throughout ageing. Nat. Ecol. Evol. 2024, 8, 1021–1034. [Google Scholar] [CrossRef] [PubMed]
  4. Weaver, R.J.; Rabinowitz, S.; Thueson, K.; Havird, J.C. Genomic Signatures of Mitonuclear Coevolution in Mammals. Mol. Biol. Evol. 2022, 39, msac233. [Google Scholar] [CrossRef]
  5. Osada, N.; Akashi, H. Mitochondrial-nuclear interactions and accelerated compensatory evolution: Evidence from the primate cytochrome C oxidase complex. Mol. Biol. Evol. 2012, 29, 337–346. [Google Scholar] [CrossRef]
  6. Lin, Y.H.; Lim, S.N.; Chen, C.Y.; Chi, H.C.; Yeh, C.T.; Lin, W.R. Functional Role of Mitochondrial DNA in Cancer Progression. Int. J. Mol. Sci. 2022, 23, 1659. [Google Scholar] [CrossRef]
  7. Sherratt, H.S. Mitochondria: Structure and function. Rev. Neurol. 1991, 147, 417–430. [Google Scholar] [PubMed]
  8. Li, X.; Wang, M.; Li, S.; Chen, Y.; Wang, M.; Wu, Z.; Sun, X.; Yao, L.; Dong, H.; Song, Y.; et al. HIF-1-induced mitochondrial ribosome protein L52: A mechanism for breast cancer cellular adaptation and metastatic initiation in response to hypoxia. Theranostics 2021, 11, 7337–7359. [Google Scholar] [CrossRef]
  9. Greber, B.J.; Ban, N. Structure and Function of the Mitochondrial Ribosome. Annu. Rev. Biochem. 2016, 85, 103–132. [Google Scholar] [CrossRef] [PubMed]
  10. Huang, G.; Li, H.; Zhang, H. Abnormal Expression of Mitochondrial Ribosomal Proteins and Their Encoding Genes with Cell Apoptosis and Diseases. Int. J. Mol. Sci. 2020, 21, 8879. [Google Scholar] [CrossRef]
  11. Greber, B.J.; Bieri, P.; Leibundgut, M.; Leitner, A.; Aebersold, R.; Boehringer, D.; Ban, N. Ribosome. The complete structure of the 55S mammalian mitochondrial ribosome. Science 2015, 348, 303–308. [Google Scholar] [CrossRef] [PubMed]
  12. Greber, B.J.; Boehringer, D.; Leitner, A.; Bieri, P.; Voigts-Hoffmann, F.; Erzberger, J.P.; Leibundgut, M.; Aebersold, R.; Ban, N. Architecture of the large subunit of the mammalian mitochondrial ribosome. Nature 2014, 505, 515–519. [Google Scholar] [CrossRef]
  13. Antonicka, H.; Shoubridge, E.A. Mitochondrial RNA Granules Are Centers for Posttranscriptional RNA Processing and Ribosome Biogenesis. Cell Rep. 2015, 10, 920–932. [Google Scholar] [CrossRef] [PubMed]
  14. Horten, P.; Colina-Tenorio, L.; Rampelt, H. Biogenesis of Mitochondrial Metabolite Carriers. Biomolecules 2020, 10, 1008. [Google Scholar] [CrossRef] [PubMed]
  15. De Silva, D.; Tu, Y.T.; Amunts, A.; Fontanesi, F.; Barrientos, A. Mitochondrial ribosome assembly in health and disease. Cell Cycle 2015, 14, 2226–2250. [Google Scholar] [CrossRef] [PubMed]
  16. O’Brien, T.W. Properties of human mitochondrial ribosomes. IUBMB Life 2003, 55, 505–513. [Google Scholar] [CrossRef]
  17. Bao, S.; Wang, X.; Li, M.; Gao, Z.; Zheng, D.; Shen, D.; Liu, L. Potential of Mitochondrial Ribosomal Genes as Cancer Biomarkers Demonstrated by Bioinformatics Results. Front. Oncol. 2022, 12, 835549. [Google Scholar] [CrossRef]
  18. Pierson, J.A.; Yang, J.E.; Wright, E.R. Recent advances in correlative cryo-light and electron microscopy. Curr. Opin. Struct. Biol. 2024, 89, 102934. [Google Scholar] [CrossRef]
  19. Khawaja, A.; Cipullo, M.; Krüger, A.; Rorbach, J. Insights into mitoribosomal biogenesis from recent structural studies. Trends Biochem. Sci. 2023, 48, 629–641. [Google Scholar] [CrossRef] [PubMed]
  20. Saurer, M.; Ramrath, D.J.F.; Niemann, M.; Calderaro, S.; Prange, C.; Mattei, S.; Scaiola, A.; Leitner, A.; Bieri, P.; Horn, E.K.; et al. Mitoribosomal small subunit biogenesis in trypanosomes involves an extensive assembly machinery. Science 2019, 365, 1144–1149. [Google Scholar] [CrossRef]
  21. Amunts, A.; Brown, A.; Toots, J.; Scheres, S.H.W.; Ramakrishnan, V. Ribosome. The structure of the human mitochondrial ribosome. Science 2015, 348, 95–98. [Google Scholar] [CrossRef] [PubMed]
  22. Aibara, S.; Singh, V.; Modelska, A.; Amunts, A. Structural basis of mitochondrial translation. Elife 2020, 9, e58362. [Google Scholar] [CrossRef]
  23. Itoh, Y.; Khawaja, A.; Laptev, I.; Cipullo, M.; Atanassov, I.; Sergiev, P.; Rorbach, J.; Amunts, A. Mechanism of mitoribosomal small subunit biogenesis and preinitiation. Nature 2022, 606, 603–608. [Google Scholar] [CrossRef] [PubMed]
  24. Harper, N.J.; Burnside, C.; Klinge, S. Principles of mitoribosomal small subunit assembly in eukaryotes. Nature 2023, 614, 175–181. [Google Scholar] [CrossRef] [PubMed]
  25. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [PubMed]
  26. Pecoraro, A.; Pagano, M.; Russo, G.; Russo, A. Ribosome Biogenesis and Cancer: Overview on Ribosomal Proteins. Int. J. Mol. Sci. 2021, 22, 5496. [Google Scholar] [CrossRef]
  27. Song, H.; Liu, H.; Wang, X.; Yang, Y.; Zhao, X.; Jiang, W.G.; Sui, L.; Song, X. Death-associated protein 3 in cancer-discrepant roles of DAP3 in tumours and molecular mechanisms. Front. Oncol. 2024, 13, 1323751. [Google Scholar] [CrossRef] [PubMed]
  28. Kissil, J.L.; Cohen, O.; Raveh, T.; Kimchi, A. Structure-function analysis of an evolutionary conserved protein, DAP3, which mediates TNF-alpha- and Fas-induced cell death. EMBO J. 1999, 18, 353–362. [Google Scholar] [CrossRef]
  29. Miyazaki, T.; Reed, J.C. A GTP-binding adapter protein couples TRAIL receptors to apoptosis-inducing proteins. Nat. Immunol. 2001, 2, 493–500. [Google Scholar] [CrossRef]
  30. Harada, T.; Iwai, A.; Miyazaki, T. Identification of DELE, a novel DAP3-binding protein which is crucial for death receptor-mediated apoptosis induction. Apoptosis 2010, 15, 1247–1255. [Google Scholar] [CrossRef] [PubMed]
  31. Takeda, S.; Iwai, A.; Nakashima, M.; Fujikura, D.; Chiba, S.; Li, H.M.; Uehara, J.; Kawaguchi, S.; Kaya, M.; Nagoya, S.; et al. LKB1 is crucial for TRAIL-mediated apoptosis induction in osteosarcoma. Anticancer. Res. 2007, 27, 761–768. [Google Scholar] [PubMed]
  32. Wazir, U.; Orakzai, M.M.; Khanzada, Z.S.; Jiang, W.G.; Sharma, A.K.; Kasem, A.; Mokbel, K. The role of death-associated protein 3 in apoptosis, anoikis and human cancer. Cancer Cell Int. 2015, 15, 39. [Google Scholar] [CrossRef] [PubMed]
  33. Liu, X.; Xi, R.; Du, X.; Wang, Y.; Cheng, L.; Yan, G.; Zhu, J.; Liu, T.; Li, F. DNA methylation of microRNA-365-1 induces apoptosis of hair follicle stem cells by targeting DAP3. Noncoding RNA Res. 2024, 9, 901–912. [Google Scholar] [CrossRef]
  34. Broecker, F.; Horton, R.; Heinrich, J.; Franz, A.; Schweiger, M.R.; Lehrach, H.; Moelling, K. The intron-enriched HERV-K(HML-10) family suppresses apoptosis, an indicator of malignant transformation. Mob. DNA 2016, 7, 25. [Google Scholar] [CrossRef] [PubMed]
  35. Miyazaki, T.; Shen, M.; Fujikura, D.; Tosa, N.; Kim, H.R.; Kon, S.; Uede, T.; Reed, J.C. Functional role of death-associated protein 3 (DAP3) in anoikis. J. Biol. Chem. 2004, 279, 44667–44672. [Google Scholar] [CrossRef] [PubMed]
  36. Yoo, Y.A.; Kim, M.J.; Park, J.K.; Chung, Y.M.; Lee, J.H.; Chi, S.G.; Kim, J.S.; Yoo, Y.D. Mitochondrial ribosomal protein L41 suppresses cell growth in association with p53 and p27Kip1. Mol. Cell Biol. 2005, 25, 6603–6616. [Google Scholar] [CrossRef] [PubMed]
  37. Guo, H.; Dong, Y.; Luo, D.; Gong, M.; Sun, J.; Wu, Z.; Liu, Z.; Zhong, L.; Jin, S. MRPL41, as a target for acupuncture, promotes neuron apoptosis in models of ischemic stroke via activating p53 pathway. Neurochem. Int. 2024, 180, 105881. [Google Scholar] [CrossRef] [PubMed]
  38. Sun, L.; Liu, Y.; Frémont, M.; Schwarz, S.; Siegmann, M.; Matthies, R.; Jost, J.P. A novel 52 kDa protein induces apoptosis and concurrently activates c-Jun N-terminal kinase 1 (JNK1) in mouse C3H10T1/2 fibroblasts. Gene 1998, 208, 157–166. [Google Scholar] [CrossRef] [PubMed]
  39. Gogvadze, V.; Zhivotovsky, B.; Orrenius, S. The Warburg effect and mitochondrial stability in cancer cells. Mol. Aspects Med. 2010, 31, 60–74. [Google Scholar] [CrossRef]
  40. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  41. Karim, L.; Kosmider, B.; Bahmed, K. Mitochondrial ribosomal stress in lung diseases. Am. J. Physiol. Lung Cell Mol. Physiol. 2022, 322, L507–L517. [Google Scholar] [CrossRef] [PubMed]
  42. Bennett, N.K.; Nakaoka, H.J.; Laurent, D.; Okimoto, R.A.; Sei, Y.; Horvai, A.E.; Bivona, T.G.; Ten Hoeve, J.; Graeber, T.G.; Nakamura, K.; et al. Primary and metastatic tumors exhibit systems-level differences in dependence on mitochondrial respiratory function. PLoS Biol. 2022, 20, e3001753. [Google Scholar] [CrossRef]
  43. Bajinka, O.; Ouedraogo, S.Y.; Golubnitschaja, O.; Li, N.; Zhan, X. Energy metabolism as the hub of advanced non-small cell lung cancer management: A comprehensive view in the framework of predictive, preventive, and personalized medicine. EPMA J. 2024, 15, 289–319. [Google Scholar] [CrossRef] [PubMed]
  44. Tang, Z.; Kang, B.; Li, C.; Chen, T.; Zhang, Z. GEPIA2: An enhanced web server for large-scale expression profiling and interactive analysis. Nucleic Acids Res. 2019, 47, W556–W560. [Google Scholar] [CrossRef] [PubMed]
  45. Ji, X.; Yang, Z.; Li, C.; Zhu, S.; Zhang, Y.; Xue, F.; Sun, S.; Fu, T.; Ding, C.; Liu, Y.; et al. Mitochondrial ribosomal protein L12 potentiates hepatocellular carcinoma by regulating mitochondrial biogenesis and metabolic reprogramming. Metabolism 2024, 152, 155761. [Google Scholar] [CrossRef] [PubMed]
  46. Ji, X.; Zhang, T.; Sun, J.; Song, X.; Ma, G.; Xu, L.; Cao, X.; Jing, Y.; Xue, F.; Zhang, W.; et al. UBASH3B-mediated MRPL12 Y60 dephosphorylation inhibits LUAD development by driving mitochondrial metabolism reprogramming. J. Exp. Clin. Cancer Res. 2024, 43, 268. [Google Scholar] [CrossRef] [PubMed]
  47. Hu, Y.; Liu, Y.; Ma, C.; Ai, K. MRPL12 Acts as A Novel Prognostic Biomarker Involved in Immune Cell Infiltration and Tumor Progression of Lung Adenocarcinoma. Int. J. Mol. Sci. 2023, 24, 2762. [Google Scholar] [CrossRef] [PubMed]
  48. Liu, Y.; Sun, H.; Li, X.; Liu, Q.; Zhao, Y.; Li, L.; Xu, B.; Hou, Y.; Jin, W. Identification of a Three-RNA Binding Proteins (RBPs) Signature Predicting Prognosis for Breast Cancer. Front. Oncol. 2021, 11, 663556. [Google Scholar] [CrossRef]
  49. Ke, H.; Dass, S.; Morrisey, J.M.; Mather, M.W.; Vaidya, A.B. The mitochondrial ribosomal protein L13 is critical for the structural and functional integrity of the mitochondrion in Plasmodium falciparum. J. Biol. Chem. 2018, 293, 8128–8137. [Google Scholar] [CrossRef]
  50. Grohmann, L.; Kitakawa, M.; Isono, K.; Goldschmidt-Reisin, S.; Graack, H.R. The yeast nuclear gene MRP-L13 codes for a protein of the large subunit of the mitochondrial ribosome. Curr. Genet. 1994, 26, 8–14. [Google Scholar] [CrossRef]
  51. Zhong, X.; He, Z.; Fan, Y.; Yin, L.; Hong, Z.; Tong, Y.; Bi, Q.; Zhu, S. Multi-omics analysis of MRPL-13 as a tumor-promoting marker from pan-cancer to lung adenocarcinoma. Aging 2023, 15, 10640–10680. [Google Scholar] [CrossRef] [PubMed]
  52. Wang, L.; Xia, Y.; Zhang, Z.; Liu, X.; Shi, J.; Wang, Y.; Li, J.; Zhnag, X.; Geng, Z.; Song, X.; et al. [High expression of MRPL13 promotes cell cycle progression and proliferation of gastric cancer cells by inhibiting p53 signaling to affect long-term prognosis]. Nan Fang Yi Ke Da Xue Xue Bao 2023, 43, 1558–1566. [Google Scholar] [CrossRef] [PubMed]
  53. Cai, M.; Li, H.; Chen, R.; Zhou, X. MRPL13 Promotes Tumor Cell Proliferation, Migration and EMT Process in Breast Cancer Through the PI3K-AKT-mTOR Pathway. Cancer Manag. Res. 2021, 13, 2009–2024. [Google Scholar] [CrossRef] [PubMed]
  54. Dai, P.; Chen, Y.; Zhang, X.; Liu, L.; Cheng, Z. MRPL13 is a metastatic and prognostic marker of breast cancer: A silico analysis accompanied with experimental validation. Gene 2025, 932, 148908. [Google Scholar] [CrossRef] [PubMed]
  55. Lee, Y.K.; Lim, J.J.; Jeoun, U.W.; Min, S.; Lee, E.B.; Kwon, S.M.; Lee, C.; Yoon, G. Lactate-mediated mitoribosomal defects impair mitochondrial oxidative phosphorylation and promote hepatoma cell invasiveness. J. Biol. Chem. 2017, 292, 20208–20217. [Google Scholar] [CrossRef] [PubMed]
  56. Xie, C.; Hu, J.; Hu, Q.; Jiang, L.; Chen, W. Classification of the mitochondrial ribosomal protein-associated molecular subtypes and identified a serological diagnostic biomarker in hepatocellular carcinoma. Front. Surg. 2023, 9, 1062659. [Google Scholar] [CrossRef] [PubMed]
  57. Li, X.Y.; He, X.Y.; Zhao, H.; Qi, L.; Lu, J.J. Identification of a novel therapeutic target for lung cancer: Mitochondrial ribosome protein L9. Pathol. Res. Pract. 2023, 248, 154625. [Google Scholar] [CrossRef]
  58. Zeng, Y.; Shi, Y.; Xu, L.; Zeng, Y.; Cui, X.; Wang, Y.; Yang, N.; Zhou, F.; Zhou, Y. Prognostic Value and Related Regulatory Networks of MRPL15 in Non-Small-Cell Lung Cancer. Front. Oncol. 2021, 11, 656172. [Google Scholar] [CrossRef] [PubMed]
  59. Wei, D.; Sun, D.; Sirera, R.; Afzal, M.Z.; Leong, T.L.; Li, X.; Wang, Y. Overexpression of MRPL19 in predicting poor prognosis and promoting the development of lung adenocarcinoma. Transl. Lung Cancer Res. 2023, 12, 1517–1538. [Google Scholar] [CrossRef]
  60. Sahu, P.K.; Salim, S.; Pp, M.; Chauhan, S.; Tomar, R.S. Reverse genetic analysis of yeast YPR099C/MRPL51 reveals a critical role of both overlapping ORFs in respiratory growth and MRPL51 in mitochondrial DNA maintenance. FEMS Yeast Res. 2019, 19, foz056. [Google Scholar] [CrossRef]
  61. Zhang, W.; Yu, L.; Xu, C.; Tang, T.; Cao, J.; Chen, L.; Pang, X.; Ren, W. MRPL51 is a downstream target of FOXM1 in promoting the malignant behaviors of lung adenocarcinoma. Oncol. Lett. 2023, 26, 298. [Google Scholar] [CrossRef] [PubMed]
  62. Gu, X.; Huang, X.; Zhang, X.; Wang, C. Development and Validation of a DNA Methylation-related Classifier of Circulating Tumour Cells to Predict Prognosis and to provide a therapeutic strategy in Lung Adenocarcinoma. Int. J. Biol. Sci. 2022, 18, 4984–5000. [Google Scholar] [CrossRef]
  63. Guttapadu, R.; Katte, T.; Sayeeram, D.; Bhatia, S.; Abraham, A.R.; Rajeev, K.; Amara, A.R.R.; Siri, S.; Bommana, K.; Rasalkar, A.A.; et al. Identification of novel biomarkers for lung squamous cell carcinoma. 3 Biotech 2023, 13, 72. [Google Scholar] [CrossRef]
  64. Han, B.; Zheng, R.; Zeng, H.; Wang, S.; Sun, K.; Chen, R.; Li, L.; Wei, W.; He, J. Cancer incidence and mortality in China, 2022. J. Natl. Cancer Cent. 2024, 4, 47–53. [Google Scholar] [CrossRef]
  65. Lin, X.; Guo, L.; Lin, X.; Wang, Y.; Zhang, G. Expression and prognosis analysis of mitochondrial ribosomal protein family in breast cancer. Sci. Rep. 2022, 12, 10658. [Google Scholar] [CrossRef]
  66. Wang, Q.; Chen, G.; Liu, L.; Peng, X.; Yang, X.; Yang, L.; Li, C. MRPS23 is a novel prognostic biomarker and promotes glioma progression. Aging 2024, 16, 2457–2474. [Google Scholar] [CrossRef] [PubMed]
  67. Liu, L.; Zhang, X.; Ding, H.; Liu, X.; Cao, D.; Liu, Y.; Liu, J.; Lin, C.; Zhang, N.; Wang, G.; et al. Arginine and lysine methylation of MRPS23 promotes breast cancer metastasis through regulating OXPHOS. Oncogene 2021, 40, 3548–3563. [Google Scholar] [CrossRef]
  68. Gatza, M.L.; Silva, G.O.; Parker, J.S.; Fan, C.; Perou, C.M. An integrated genomics approach identifies drivers of proliferation in luminal-subtype human breast cancer. Nat. Genet. 2014, 46, 1051–1059. [Google Scholar] [CrossRef] [PubMed]
  69. Klæstad, E.; Opdahl, S.; Engstrøm, M.J.; Ytterhus, B.; Wik, E.; Bofin, A.M.; Valla, M. MRPS23 amplification and gene expression in breast cancer; association with proliferation and the non-basal subtypes. Breast Cancer Res. Treat. 2020, 180, 73–86. [Google Scholar] [CrossRef]
  70. Gao, Y.; Li, F.; Zhou, H.; Yang, Y.; Wu, R.; Chen, Y.; Li, W.; Li, Y.; Xu, X.; Ke, C.; et al. Down-regulation of MRPS23 inhibits rat breast cancer proliferation and metastasis. Oncotarget 2017, 8, 71772–71781. [Google Scholar] [CrossRef]
  71. Min, L.; Chen, L.; Huang, D.; Zhang, Y.; You, A.; Yan, X.; Li, Z.H. LncRNA HIF1A-AS2 promotes triple-negative breast cancer progression and paclitaxel resistance via MRPS23 protein. Heliyon 2024, 10, e36469. [Google Scholar] [CrossRef]
  72. Oviya, R.P.; Thangaretnam, K.P.; Ramachandran, B.; Ramanathan, P.; Jayavelu, S.; Gopal, G.; Rajkumar, T. Mitochondrial ribosomal small subunit (MRPS) MRPS23 protein-protein interaction reveals phosphorylation by CDK11-p58 affecting cell proliferation and knockdown of MRPS23 sensitizes breast cancer cells to CDK1 inhibitors. Mol. Biol. Rep. 2022, 49, 9521–9534. [Google Scholar] [CrossRef]
  73. Pu, M.; Wang, J.; Huang, Q.; Zhao, G.; Xia, C.; Shang, R.; Zhang, Z.; Bian, Z.; Yang, X.; Tao, K. High MRPS23 expression contributes to hepatocellular carcinoma proliferation and indicates poor survival outcomes. Tumour Biol. 2017, 39, 1010428317709127. [Google Scholar] [CrossRef]
  74. Lyng, H.; Brøvig, R.S.; Svendsrud, D.H.; Holm, R.; Kaalhus, O.; Knutstad, K.; Oksefjell, H.; Sundfør, K.; Kristensen, G.B.; Stokke, T. Gene expressions and copy numbers associated with metastatic phenotypes of uterine cervical cancer. BMC Genom. 2006, 7, 268. [Google Scholar] [CrossRef]
  75. Davies, S.M.; Lopez Sanchez, M.I.; Narsai, R.; Shearwood, A.M.; Razif, M.F.; Small, I.D.; Whelan, J.; Rackham, O.; Filipovska, A. MRPS27 is a pentatricopeptide repeat domain protein required for the translation of mitochondrially encoded proteins. FEBS Lett. 2012, 586, 3555–3561. [Google Scholar] [CrossRef]
  76. Zheng, C.; Yao, H.; Lu, L.; Li, H.; Zhou, L.; He, X.; Xu, X.; Xia, H.; Ding, S.; Yang, Y.; et al. Dysregulated Ribosome Biogenesis Is a Targetable Vulnerability in Triple-Negative Breast Cancer: MRPS27 as a Key Mediator of the Stemness-inhibitory Effect of Lovastatin. Int. J. Biol. Sci. 2024, 20, 2130–2148. [Google Scholar] [CrossRef] [PubMed]
  77. Vishnubalaji, R.; Alajez, N.M. Transcriptional landscape associated with TNBC resistance to neoadjuvant chemotherapy revealed by single-cell RNA-seq. Mol. Ther. Oncolyt. 2021, 23, 151–162. [Google Scholar] [CrossRef]
  78. Wang, X.; Zhang, L.; Chen, Z.; Ma, Y.; Zhao, Y.; Rewuti, A.; Zhang, F.; Fu, D.; Han, Y. Association between 5p12 genomic markers and breast cancer susceptibility: Evidence from 19 case-control studies. PLoS ONE 2013, 8, e73611. [Google Scholar] [CrossRef] [PubMed]
  79. Ghoussaini, M.; French, J.D.; Michailidou, K.; Nord, S.; Beesley, J.; Canisus, S.; Hillman, K.M.; Kaufmann, S.; Sivakumaran, H.; Moradi Marjaneh, M.; et al. Evidence that the 5p12 Variant rs10941679 Confers Susceptibility to Estrogen-Receptor-Positive Breast Cancer through FGF10 and MRPS30 Regulation. Am. J. Hum. Genet. 2016, 99, 903–911. [Google Scholar] [CrossRef]
  80. Quigley, D.A.; Fiorito, E.; Nord, S.; Van Loo, P.; Alnæs, G.G.; Fleischer, T.; Tost, J.; Moen Vollan, H.K.; Tramm, T.; Overgaard, J.; et al. The 5p12 breast cancer susceptibility locus affects MRPS30 expression in estrogen-receptor positive tumors. Mol. Oncol. 2014, 8, 273–284. [Google Scholar] [CrossRef] [PubMed]
  81. Shirani, N.; Mahdi-Esferizi, R.; Eshraghi Samani, R.; Tahmasebian, S.; Yaghoobi, H. In silico identification and in vitro evaluation of MRPS30-DT lncRNA and MRPS30 gene expression in breast cancer. Cancer Rep. 2024, 7, e2114. [Google Scholar] [CrossRef] [PubMed]
  82. Grigoriadis, A.; Mackay, A.; Reis-Filho, J.S.; Steele, D.; Iseli, C.; Stevenson, B.J.; Jongeneel, C.V.; Valgeirsson, H.; Fenwick, K.; Iravani, M.; et al. Establishment of the epithelial-specific transcriptome of normal and malignant human breast cells based on MPSS and array expression data. Breast Cancer Res. 2006, 8, R56. [Google Scholar] [CrossRef]
  83. Zhou, W.; Ouyang, J.; Li, J.; Liu, F.; An, T.; Cheng, L.; Kuo, Z.C.; Zhang, C.; He, Y. MRPS17 promotes invasion and metastasis through PI3K/AKT signal pathway and could be potential prognostic marker for gastric cancer. J. Cancer 2021, 12, 4849–4861. [Google Scholar] [CrossRef]
  84. Buffa, F.M.; Harris, A.L.; West, C.M.; Miller, C.J. Large meta-analysis of multiple cancers reveals a common, compact and highly prognostic hypoxia metagene. Br. J. Cancer 2010, 102, 428–435. [Google Scholar] [CrossRef] [PubMed]
  85. Gong, P.J.; Shao, Y.C.; Huang, S.R.; Zeng, Y.F.; Yuan, X.N.; Xu, J.J.; Yin, W.N.; Wei, L.; Zhang, J.W. Hypoxia-Associated Prognostic Markers and Competing Endogenous RNA Co-Expression Networks in Breast Cancer. Front. Oncol. 2020, 10, 579868. [Google Scholar] [CrossRef] [PubMed]
  86. Sotgia, F.; Whitaker-Menezes, D.; Martinez-Outschoorn, U.E.; Salem, A.F.; Tsirigos, A.; Lamb, R.; Sneddon, S.; Hulit, J.; Howell, A.; Lisanti, M.P. Mitochondria “fuel” breast cancer metabolism: Fifteen markers of mitochondrial biogenesis label epithelial cancer cells, but are excluded from adjacent stromal cells. Cell Cycle 2012, 11, 4390–4401. [Google Scholar] [CrossRef]
  87. Li, J.; Zhou, W.; Wei, J.; Xiao, X.; An, T.; Wu, W.; He, Y. Prognostic Value and Biological Functions of RNA Binding Proteins in Stomach Adenocarcinoma. Onco Targets Ther. 2021, 14, 1689–1705. [Google Scholar] [CrossRef] [PubMed]
  88. Hu, J.L.; Huang, M.J.; Halina, H.; Qiao, K.; Wang, Z.Y.; Lu, J.J.; Yin, C.L.; Gao, F. Identification of a novel inflammatory-related gene signature to evaluate the prognosis of gastric cancer patients. World J. Gastrointest. Oncol. 2024, 16, 945–967. [Google Scholar] [CrossRef]
  89. Box, J.M.; Anderson, J.M.; Stuart, R.A. Mutation of the PEBP-like domain of the mitoribosomal MrpL35/mL38 protein results in production of nascent chains with impaired capacity to assemble into OXPHOS complexes. Mol. Biol. Cell 2023, 34, ar131. [Google Scholar] [CrossRef] [PubMed]
  90. Box, J.M.; Kaur, J.; Stuart, R.A. MrpL35, a mitospecific component of mitoribosomes, plays a key role in cytochrome c oxidase assembly. Mol. Biol. Cell 2017, 28, 3489–3499. [Google Scholar] [CrossRef] [PubMed]
  91. Zhang, L.; Lu, P.; Yan, L.; Yang, L.; Wang, Y.; Chen, J.; Dai, J.; Li, Y.; Kang, Z.; Bai, T.; et al. MRPL35 Is Up-Regulated in Colorectal Cancer and Regulates Colorectal Cancer Cell Growth and Apoptosis. Am. J. Pathol. 2019, 189, 1105–1120. [Google Scholar] [CrossRef]
  92. Yuan, L.; Li, J.X.; Yang, Y.; Chen, Y.; Ma, T.T.; Liang, S.; Bu, Y.; Yu, L.; Nan, Y. Depletion of MRPL35 inhibits gastric carcinoma cell proliferation by regulating downstream signaling proteins. World J. Gastroenterol. 2021, 27, 1785–1804. [Google Scholar] [CrossRef]
  93. Hou, W.; Chen, J.; Wang, Y. MRPL35 Induces Proliferation, Invasion, and Glutamine Metabolism in NSCLC Cells by Upregulating SLC7A5 Expression. Clin. Respir. J. 2024, 18, e13799. [Google Scholar] [CrossRef]
  94. Zhao, C.; Chen, L.; Jin, Z.; Liu, H.; Ma, C.; Zhou, H.; Xu, L.; Zhou, S.; Shi, Y.; Li, W.; et al. Knockdown of MRPL35 promotes cell apoptosis and inhibits cell proliferation in non-small-cell lung cancer. BMC Pulm. Med. 2023, 23, 507. [Google Scholar] [CrossRef] [PubMed]
  95. Zhang, Z.; Sun, J.; Jin, C.; Zhang, L.; Wu, L.; Tian, G. Identification and validation of a fatty acid metabolism gene signature for the promotion of metastasis in liver cancer. Oncol. Lett. 2023, 26, 457. [Google Scholar] [CrossRef] [PubMed]
  96. Yuan, L.; Yang, Y.; Li, X.; Zhou, X.; Du, Y.H.; Liu, W.J.; Zhang, L.; Yu, L.; Ma, T.T.; Li, J.X.; et al. 18β-glycyrrhetinic acid regulates mitochondrial ribosomal protein L35-associated apoptosis signaling pathways to inhibit proliferation of gastric carcinoma cells. World J. Gastroenterol. 2022, 28, 2437–2456. [Google Scholar] [CrossRef] [PubMed]
  97. Wang, A.; Zhang, Q.; Zhang, D.; Xiu, Y.; Ding, Y.; Liu, L. Effect of silencing mitochondrial ribosomal protein L35 gene on growth of human esophageal cancer TE-1 cells. J. Jilin Univ. 2019, 45, 28–32. [Google Scholar]
  98. Sotgia, F.; Lisanti, M.P. Mitochondrial biomarkers predict tumor progression and poor overall survival in gastric cancers: Companion diagnostics for personalized medicine. Oncotarget 2017, 8, 67117–67128. [Google Scholar] [CrossRef] [PubMed]
  99. Wang, C.; Chen, Y.; Zhou, R.; Yang, Y.; Fang, Y. Systematic Analysis of Tumor Stem Cell-related Gene Characteristics to Predict the PD-L1 Immunotherapy and Prognosis of Gastric Cancer. Curr. Med. Chem. 2024, 31, 2467–2482. [Google Scholar] [CrossRef]
  100. Zhao, J.W.; Zhao, W.Y.; Cui, X.H.; Xing, L.; Shi, J.C.; Yu, L. The role of the mitochondrial ribosomal protein family in detecting hepatocellular carcinoma and predicting prognosis, immune features, and drug sensitivity. Clin. Transl. Oncol. 2024, 26, 496–514. [Google Scholar] [CrossRef] [PubMed]
  101. Ma, T.; Huang, Y.B.; Chen, J.; Zhang, L.; Liu, Y.H.; Lu, C.H. MRPL21 promotes HCC proliferation through TP53 mutation-induced apoptotic resistance. Tissue Cell 2024, 86, 102298. [Google Scholar] [CrossRef]
  102. Shi, Z.Z.; Shang, L.; Jiang, Y.Y.; Shi, F.; Xu, X.; Wang, M.R.; Hao, J.J. Identification of genomic biomarkers associated with the clinicopathological parameters and prognosis of esophageal squamous cell carcinoma. Cancer Biomark. 2015, 15, 755–761. [Google Scholar] [CrossRef]
  103. Aasebø, E.; Berven, F.S.; Hovland, R.; Døskeland, S.O.; Bruserud, Ø.; Selheim, F.; Hernandez-Valladares, M. The Progression of Acute Myeloid Leukemia from First Diagnosis to Chemoresistant Relapse: A Comparison of Proteomic and Phosphoproteomic Profiles. Cancers 2020, 12, 1466. [Google Scholar] [CrossRef]
  104. Khashei Varnamkhasti, K.; Moghanibashi, M.; Naeimi, S. Genes whose expressions in the primary lung squamous cell carcinoma are able to accurately predict the progression of metastasis through lymphatic system, inferred from a bioinformatics analyses. Sci. Rep. 2023, 13, 6733. [Google Scholar] [CrossRef] [PubMed]
  105. Lin, Y.X.; Pan, J.Y.; Feng, W.D.; Huang, T.C.; Li, C.Z. MRPL48 is a novel prognostic and predictive biomarker of hepatocellular carcinoma. Eur. J. Med. Res. 2023, 28, 589. [Google Scholar] [CrossRef]
  106. Hu, T.T.; Yang, J.W.; Yan, Y.; Chen, Y.Y.; Xue, H.B.; Xiang, Y.Q.; Ye, L.C. Detection of genes responsible for cetuximab sensitization in colorectal cancer cells using CRISPR-Cas9. Biosci. Rep. 2020, 40, BSR20201125. [Google Scholar] [CrossRef] [PubMed]
  107. Liu, L.; Luo, C.; Luo, Y.; Chen, L.; Liu, Y.; Wang, Y.; Han, J.; Zhang, Y.; Wei, N.; Xie, Z.; et al. MRPL33 and its splicing regulator hnRNPK are required for mitochondria function and implicated in tumor progression. Oncogene 2018, 37, 86–94. [Google Scholar] [CrossRef]
  108. Li, J.; Feng, D.; Gao, C.; Zhang, Y.; Xu, J.; Wu, M.; Zhan, X. Isoforms S and L of MRPL33 from alternative splicing have isoform-specific roles in the chemoresponse to epirubicin in gastric cancer cells via the PI3K/AKT signaling pathway. Int. J. Oncol. 2019, 54, 1591–1600. [Google Scholar] [CrossRef]
  109. Zhu, X.; Liu, Y.; Xu, J.; Cheng, Z.; Yu, Y.; Chu, M.; Lu, X.; Yuan, W. miR-608 rs4919510 Polymorphism May Affect Susceptibility to Colorectal Cancer by Upregulating MRPL43 Expression. DNA Cell Biol. 2020, 39, 2017–2027. [Google Scholar] [CrossRef] [PubMed]
  110. Abdul Aziz, N.A.; Mokhtar, N.M.; Harun, R.; Mollah, M.M.; Mohamed Rose, I.; Sagap, I.; Mohd Tamil, A.; Wan Ngah, W.Z.; Jamal, R. A 19-Gene expression signature as a predictor of survival in colorectal cancer. BMC Med. Genom. 2016, 9, 58. [Google Scholar] [CrossRef]
  111. Xin, C.; Lai, Y.; Ji, L.; Wang, Y.; Li, S.; Hao, L.; Zhang, W.; Meng, R.; Xu, J.; Hong, Y.; et al. A novel 9-gene signature for the prediction of postoperative recurrence in stage II/III colorectal cancer. Front. Genet. 2023, 13, 1097234. [Google Scholar] [CrossRef]
  112. Chen, K.; He, Y.; Liu, Y.; Yang, X. Gene signature associated with neuro-endocrine activity predicting prognosis of pancreatic carcinoma. Mol. Genet. Genomic Med. 2019, 7, e00729. [Google Scholar] [CrossRef] [PubMed]
  113. Liu, Z.; Zeinalzadeh, Z.; Huang, T.; Han, Y.; Peng, L.; Wang, D.; Zhou, Z.; Ousmane, D.; Wang, J. Mitochondria-related chemoradiotherapy resistance genes-based machine learning model associated with immune cell infiltration on the prognosis of esophageal cancer and its value in pan-cancer. Transl. Oncol. 2024, 42, 101896. [Google Scholar] [CrossRef] [PubMed]
  114. Tian, A.; Pu, K.; Li, B.; Li, M.; Liu, X.; Gao, L.; Mao, X. Weighted gene coexpression network analysis reveals hub genes involved in cholangiocarcinoma progression and prognosis. Hepatol. Res. 2019, 49, 1195–1206. [Google Scholar] [CrossRef] [PubMed]
  115. Kim, D.W.; Kim, Y.C.; Kovari, B.P.; Martinez, M.; Miao, R.; Yu, J.; Mehta, R.; Strosberg, J.; Imanirad, I.; Kim, R.D. Biomarker Analysis from a Phase I/Ib Study of Regorafenib and Nivolumab in Mismatch Repair-Proficient Advanced Refractory Colorectal Cancer. Cancers 2024, 16, 556. [Google Scholar] [CrossRef] [PubMed]
  116. Zhou, C.; Chen, Z.; Peng, C.; Chen, C.; Li, H. Long Noncoding RNA TRIM52-AS1 Sponges miR-514a-5p to Facilitate Hepatocellular Carcinoma Progression Through Increasing MRPS18A. Cancer Biother. Radiopharm. 2021, 36, 211–219. [Google Scholar] [CrossRef]
  117. Sørensen, K.M.; Meldgaard, T.; Melchjorsen, C.J.; Fridriksdottir, A.J.; Pedersen, H.; Petersen, O.W.; Kristensen, P. Upregulation of Mrps18a in breast cancer identified by selecting phage antibody libraries on breast tissue sections. BMC Cancer 2017, 17, 19. [Google Scholar] [CrossRef] [PubMed]
  118. Zhuang, L.; Meng, Z.; Yang, Z. MRPL27 contributes to unfavorable overall survival and disease-free survival from cholangiocarcinoma patients. Int. J. Med. Sci. 2021, 18, 936–943. [Google Scholar] [CrossRef] [PubMed]
  119. Huang, L.; Xu, D.; Qian, Y.; Zhang, X.; Guo, H.; Sha, M.; Hu, R.; Kong, X.; Xia, Q.; Zhang, Y. A gene signature is critical for intrahepatic cholangiocarcinoma stem cell self-renewal and chemotherapeutic response. Stem Cell Res. Ther. 2022, 13, 292. [Google Scholar] [CrossRef]
  120. Zhang, H.M.; Li, Z.Y.; Dai, Z.T.; Wang, J.; Li, L.W.; Zong, Q.B.; Li, J.P.; Zhang, T.C.; Liao, X.H. Interaction of MRPL9 and GGCT Promotes Cell Proliferation and Migration by Activating the MAPK/ERK Pathway in Papillary Thyroid Cancer. Int. J. Mol. Sci. 2022, 23, 11989. [Google Scholar] [CrossRef]
  121. Xu, H.; Zou, R.; Li, F.; Liu, J.; Luan, N.; Wang, S.; Zhu, L. MRPL15 is a novel prognostic biomarker and therapeutic target for epithelial ovarian cancer. Cancer Med. 2021, 10, 3655–3673. [Google Scholar] [CrossRef] [PubMed]
  122. Zhao, L.; Tang, Y.; Yang, J.; Lin, F.; Liu, X.; Zhang, Y.; Chen, J. Integrative analysis of circadian clock with prognostic and immunological biomarker identification in ovarian cancer. Front. Mol. Biosci. 2023, 10, 1208132. [Google Scholar] [CrossRef]
  123. Zou, X.; Hu, X.; He, F.; Zhang, M.; Kong, X.; Rui, S.; Liu, Y.; Wang, L.; Zheng, X.; Liu, J.; et al. LncRNA LINC00152 promotes oral squamous cell carcinoma growth via enhancing Upstream Transcription Factor 1 mediated Mitochondrial Ribosomal Protein L52 transcription. J. Oral. Pathol. Med. 2022, 51, 454–463. [Google Scholar] [CrossRef]
  124. Guo, T.; Zambo, K.D.A.; Zamuner, F.T.; Ou, T.; Hopkins, C.; Kelley, D.Z.; Wulf, H.A.; Winkler, E.; Erbe, R.; Danilova, L.; et al. Chromatin structure regulates cancer-specific alternative splicing events in primary HPV-related oropharyngeal squamous cell carcinoma. Epigenetics 2020, 15, 959–971. [Google Scholar] [CrossRef]
  125. Wei, Z.; Liu, C.; Liang, J.; Zhou, X.; Xue, K.; Wang, K.; Zhang, X. Characterization of Mitoribosomal Small Subunit unit genes related immune and pharmacogenomic landscapes in renal cell carcinoma. IUBMB Life 2024, 76, 647–665. [Google Scholar] [CrossRef] [PubMed]
  126. Yang, X.; Han, B.; Xie, Q.; Li, Y.; Li, Q.; Hu, X.; Zhao, H.; Xu, X. Low Expression of Mitochondrial Ribosomal Protein S5 is Associated With Poor Prognosis in Patients With Clear Cell Renal Cell Carcinoma. Appl. Immunohistochem. Mol. Morphol. 2025, 33, 22–28. [Google Scholar] [CrossRef] [PubMed]
  127. Hu, G.; Jiang, Q.; Liu, L.; Peng, H.; Wang, Y.; Li, S.; Tang, Y.; Yu, J.; Yang, J.; Liu, Z. Integrated Analysis of RNA-Binding Proteins Associated With the Prognosis and Immunosuppression in Squamous Cell Carcinoma of Head and Neck. Front. Genet. 2020, 11, 571403. [Google Scholar] [CrossRef]
  128. Abaji, R.; Ceppi, F.; Patel, S.; Gagné, V.; Xu, C.J.; Spinella, J.F.; Colombini, A.; Parasole, R.; Buldini, B.; Basso, G.; et al. Genetic risk factors for VIPN in childhood acute lymphoblastic leukemia patients identified using whole-exome sequencing. Pharmacogenomics 2018, 19, 1181–1193. [Google Scholar] [CrossRef]
  129. Liu, X.; Zhuang, H.; Li, F.; Lu, Y.; Pei, R. Prognostic Significance of Dual-Specificity Phosphatase 23 Expression in Acute Myeloid Leukemia. J. Blood Med. 2024, 15, 35–50. [Google Scholar] [CrossRef] [PubMed]
  130. Fu, Q.; Hong, R.; Zhou, H.; Li, Y.; Liu, X.; Gong, J.; Wang, X.; Chen, J.; Ran, H.; Wang, L.; et al. Proteomics reveals MRPL4 as a high-risk factor and a potential diagnostic biomarker for prostate cancer. Proteomics 2022, 22, e2200081. [Google Scholar] [CrossRef] [PubMed]
  131. Menyhárt, O.; Fekete, J.T.; Győrffy, B. Gene expression-based biomarkers designating glioblastomas resistant to multiple treatment strategies. Carcinogenesis 2021, 42, 804–813. [Google Scholar] [CrossRef] [PubMed]
  132. Liu, J.; Wu, S.; Xie, X.; Wang, Z.; Lei, Q. Identification of potential crucial genes and key pathways in osteosarcoma. Hereditas 2020, 157, 29. [Google Scholar] [CrossRef]
  133. Bell, J.L.; Hagemann, S.; Holien, J.K.; Liu, T.; Nagy, Z.; Schulte, J.H.; Misiak, D.; Hüttelmaier, S. Identification of RNA-Binding Proteins as Targetable Putative Oncogenes in Neuroblastoma. Int. J. Mol. Sci. 2020, 21, 5098. [Google Scholar] [CrossRef]
  134. Bae, J.S.; Lee, J.W.; Yoo, J.E.; Joung, J.G.; Yoo, K.H.; Koo, H.H.; Song, Y.M.; Sung, K.W. Genome-Wide Association Study for the Identification of Novel Genetic Variants Associated with the Risk of Neuroblastoma in Korean Children. Cancer Res. Treat. 2020, 52, 1251–1261. [Google Scholar] [CrossRef] [PubMed]
  135. Revathi Paramasivam, O.; Gopisetty, G.; Subramani, J.; Thangarajan, R. Expression and affinity purification of recombinant mammalian mitochondrial ribosomal small subunit (MRPS) proteins and protein-protein interaction analysis indicate putative role in tumourigenic cellular processes. J. Biochem. 2021, 169, 675–692. [Google Scholar] [CrossRef]
  136. Song, Y.; Li, G.; Zhang, Z.; Liu, Y.; Jia, H.; Zhang, C.; Wang, J.; Hu, Y.; Hao, F.; Liu, X.; et al. Integrated Transcriptomic Landscape and Deep Learning Based Survival Prediction in Uterine Sarcomas. Cancer Res. Treat. 2024. [Google Scholar] [CrossRef] [PubMed]
  137. Kim, H.J.; Maiti, P.; Barrientos, A. Mitochondrial ribosomes in cancer. Semin. Cancer Biol. 2017, 47, 67–81. [Google Scholar] [CrossRef] [PubMed]
  138. Zong, Y.; Li, H.; Liao, P.; Chen, L.; Pan, Y.; Zheng, Y.; Zhang, C.; Liu, D.; Zheng, M.; Gao, J. Mitochondrial dysfunction: Mechanisms and advances in therapy. Signal Transduct. Target. Ther. 2024, 9, 124. [Google Scholar] [CrossRef]
  139. Berner, M.J.; Wall, S.W.; Echeverria, G.V. Deregulation of mitochondrial gene expression in cancer: Mechanisms and therapeutic opportunities. Br. J. Cancer 2024, 131, 1415–1424. [Google Scholar] [CrossRef]
  140. Echtenkamp, F.J.; Ishida, R.; Rivera-Marquez, G.M.; Maisiak, M.; Johnson, O.T.; Shrimp, J.H.; Sinha, A.; Ralph, S.J.; Nisbet, I.; Cherukuri, M.K.; et al. Mitoribosome sensitivity to HSP70 inhibition uncovers metabolic liabilities of castration-resistant prostate cancer. PNAS Nexus 2023, 2, pgad115. [Google Scholar] [CrossRef]
  141. Vendramin, R.; Verheyden, Y.; Ishikawa, H.; Goedert, L.; Nicolas, E.; Saraf, K.; Armaos, A.; Delli Ponti, R.; Izumikawa, K.; Mestdagh, P.; et al. SAMMSON fosters cancer cell fitness by concertedly enhancing mitochondrial and cytosolic translation. Nat. Struct. Mol. Biol. 2018, 25, 1035–1046. [Google Scholar] [CrossRef]
  142. Dewaele, S.; Delhaye, L.; De Paepe, B.; de Bony, E.J.; De Wilde, J.; Vanderheyden, K.; Anckaert, J.; Yigit, N.; Nuytens, J.; Vanden Eynde, E.; et al. The long non-coding RNA SAMMSON is essential for uveal melanoma cell survival. Oncogene 2022, 41, 15–25. [Google Scholar] [CrossRef]
  143. Dijk, S.N.; Protasoni, M.; Elpidorou, M.; Kroon, A.M.; Taanman, J.W. Mitochondria as target to inhibit proliferation and induce apoptosis of cancer cells: The effects of doxycycline and gemcitabine. Sci. Rep. 2020, 10, 4363. [Google Scholar] [CrossRef] [PubMed]
  144. Protasoni, M.; Kroon, A.M.; Taanman, J.W. Mitochondria as oncotarget: A comparison between the tetracycline analogs doxycycline and COL-3. Oncotarget 2018, 9, 33818–33831. [Google Scholar] [CrossRef]
  145. Lamb, R.; Ozsvari, B.; Lisanti, C.L.; Tanowitz, H.B.; Howell, A.; Martinez-Outschoorn, U.E.; Sotgia, F.; Lisanti, M.P. Antibiotics that target mitochondria effectively eradicate cancer stem cells, across multiple tumor types: Treating cancer like an infectious disease. Oncotarget 2015, 6, 4569–4584. [Google Scholar] [CrossRef]
  146. Koch, D.T.; Yu, H.; Beirith, I.; Schirren, M.; Drefs, M.; Liu, Y.; Knoblauch, M.; Koliogiannis, D.; Sheng, W.; De Toni, E.N.; et al. Tigecycline causes loss of cell viability mediated by mitochondrial OXPHOS and RAC1 in hepatocellular carcinoma cells. J. Transl. Med. 2023, 21, 876. [Google Scholar] [CrossRef] [PubMed]
  147. Huang, S.W.; Sun, M.T.; Lee, W.S.; Su, Y.S.; Lee, Y.T.; Chiang, M.H.; Wang, Y.C.; Yang, Y.S.; Tzeng, S.C.; Wu, T.Y.; et al. Cancer as an infectious disease: A different treatment alternative using a combination of tigecycline and pyrvinium pamoate—An example of breast cancer. J. Microbiol. Immunol. Infect. 2022, 55, 51–59. [Google Scholar] [CrossRef] [PubMed]
  148. Hu, B.; Guo, Y. Inhibition of mitochondrial translation as a therapeutic strategy for human ovarian cancer to overcome chemoresistance. Biochem. Biophys. Res. Commun. 2019, 509, 373–378. [Google Scholar] [CrossRef]
  149. Delaunay, S.; Pascual, G.; Feng, B.; Klann, K.; Behm, M.; Hotz-Wagenblatt, A.; Richter, K.; Zaoui, K.; Herpel, E.; Münch, C.; et al. Mitochondrial RNA modifications shape metabolic plasticity in metastasis. Nature 2022, 607, 593–603. [Google Scholar] [CrossRef]
  150. Reed, G.A.; Schiller, G.J.; Kambhampati, S.; Tallman, M.S.; Douer, D.; Minden, M.D.; Yee, K.W.; Gupta, V.; Brandwein, J.; Jitkova, Y.; et al. A Phase 1 study of intravenous infusions of tigecycline in patients with acute myeloid leukemia. Cancer Med. 2016, 5, 3031–3040. [Google Scholar] [CrossRef]
  151. Scatena, C.; Roncella, M.; Di Paolo, A.; Aretini, P.; Menicagli, M.; Fanelli, G.; Marini, C.; Mazzanti, C.M.; Ghilli, M.; Sotgia, F.; et al. Doxycycline, an Inhibitor of Mitochondrial Biogenesis, Effectively Reduces Cancer Stem Cells (CSCs) in Early Breast Cancer Patients: A Clinical Pilot Study. Front. Oncol. 2018, 8, 452. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Mitoribosome structure and biogenesis.
Figure 1. Mitoribosome structure and biogenesis.
Medicina 61 00096 g001
Figure 2. The roles of DAP3, MRPL41, and MRPS30 in apoptosis.
Figure 2. The roles of DAP3, MRPL41, and MRPS30 in apoptosis.
Medicina 61 00096 g002
Figure 3. Differentially expressed MRPs in relation to lung cancer. (a) Eight upregulated differentially expressed MRPs in LUADs; (b) 22 upregulated and one downregulated differentially expressed MRPs in LUSCs. |log2 (FC)|> 1, q value < 0.01.
Figure 3. Differentially expressed MRPs in relation to lung cancer. (a) Eight upregulated differentially expressed MRPs in LUADs; (b) 22 upregulated and one downregulated differentially expressed MRPs in LUSCs. |log2 (FC)|> 1, q value < 0.01.
Medicina 61 00096 g003
Figure 4. Differentially expressed MRPs in digestive system cancers. (a) Gastric cancer; (b) COAD; (c) READ; (d) ESCA. |log2 (FC)| > 1, q value < 0.01.
Figure 4. Differentially expressed MRPs in digestive system cancers. (a) Gastric cancer; (b) COAD; (c) READ; (d) ESCA. |log2 (FC)| > 1, q value < 0.01.
Medicina 61 00096 g004
Table 1. Differentially expressed MRPs across subtypes of breast cancer in TCGA.
Table 1. Differentially expressed MRPs across subtypes of breast cancer in TCGA.
SubtypesSamplesExpression PatternDifferentially Expressed MRPs *
Luminal A563UpregulationMRPL12, MRPL14, MRPL24, MRPL41, MRPL55, MRPS12, MRPS30, MRPS34
Luminal B210UpregulationMRPL12, MRPL13, MRPL14, MRPL15, MRPL17, MRPL21, MRPL24, MRPL27, MRPL47, MRPL55, MRPL58, CHCHD1, DAP3, MRPS12, MRPS23, MRPS28, MRPS30, MRPS34
Basal-like190UpregulationGADD45GIP1, MRPL2, MRPL4, MRPL9, MRPL12, MRPL13, MRPL14, MRPL15, MRPL36, MRPL37, MRPL47, MRPL51, MRPL55, DAP3, MRPS12, MRPS21, MRPS34
HER2-enriched79UpregulationMRPL12, MRPL13, MRPL14, MRPL27, MRPL41, MRPL47, MRPL58, CHCHD1, MRPS12, MRPS34
Normal-like70UpregulationMRPL13
* |log2 (FC)|> 1, p value < 0.001.
Table 2. Differentially expressed MRPs across pan-cancer based on the GEPIA database.
Table 2. Differentially expressed MRPs across pan-cancer based on the GEPIA database.
CancerExpression
Pattern
Differentially Expressed MRPs *
AMLDownregulationMRPL2, MRPL4, MRPL10, MRPL11, MRPL12, MRPL13, MRPL14, MRPL16, MRPL17, MRPL18, MRPL19, MRPL20, MRPL21, MRPL22, MRPL27, MRPL28, MRPL30, MRPL32, MRPL35, MRPL36, MRPL37, MRPL39, MRPL40, MRPL41, MRPL42, MRPL46, MRPL47, MRPL48, MRPL49, MRPL50, MRPL51, MRPL52, MRPL55, GADD45GIP1, MRPS18A, MRPS18B, MRPS18C, MRPS2, MRPS6, MRPS7, MRPS9, MRPS10, MRPS12, MRPS14, MRPS15, MRPS16, MRPS17, MRPS21, MRPS22, MRPS23, MRPS24, MRPS28, DAP3, MRPS33, MRPS34, MRPS35, CHCHD1, AURKAIP1
Breast invasive carcinomaUpregulationMRPL3, MRPL13, MRPL14, MRPL17, MRPL24, MRPL42, MRPL47, MRPS30, MRPS23, DAP3, MRPS34, MRPS35
Glioblastoma multiformeUpregulationMRPL3, MRPL11, MRPL13, MRPL14, MRPL19, MRPL22, MRPL24, MRPL36, MRPL39, MRPL40, MRPL42, MRPL44, MRPL47, MRPL49, MRPL52, MRPL53, MRPL54, MRPS6, MRPS10, MRPS12, MRPS14, MRPS16, MRPS23, MRPS28, DAP3, MRPS33, MRPS35
DownregulationMRPL41
Head and neck squamous cell carcinomaUpregulationMRPL47
HCCUpregulationMRPL9, MRPL13, MRPL14, MRPL15, MRPL17, MRPL21, MRPL24, MRPL47, MRPL55, MRPS16, MRPS17, MRPS21, MRPS23, DAP3
Kidney renal clear cell carcinomaDownregulationMRPS25
Lower grade gliomaUpregulationMRPL3, MRPL11, MRPL14, MRPL19, MRPL24, MRPL42, MRPS18B, MRPS14, MRPS16, MRPS23, MRPS28
DownregulationMRPL41
Ovarian serous cystadenocarcinomaUpregulationMRPL3, MRPL12, MRPL13, MRPL14, MRPL15, MRPL34, MRPL35, MRPL37, MRPL47, MRPL48, MRPS18A, MRPS11, MRPS12, MRPS15, MRPS16, MRPS35, AURKAIP1
Pancreatic adenocarcinomaUpregulationMRPL1, MRPL3, MRPL4, MRPL9, MRPL11, MRPL12, MRPL13, MRPL14, MRPL15, MRPL16, MRPL17, MRPL18, MRPL19, MRPL20, MRPL21, MRPL22, MRPL24, MRPL27, MRPL28, MRPL30, MRPL33, MRPL34, MRPL35, MRPL36, MRPL41, MRPL42, MRPL43, MRPL44, MRPL46, MRPL47, MRPL48, MRPL49, MRPL50, MRPL51, MRPL52, MRPL53, MRPL54, MRPL55, MRPL57, GADD45GIP1, MRPS30, MRPS18A, MRPS18C, MRPS5, MRPS6, MRPS7, MRPS10, MRPS11, MRPS12, MRPS14, MRPS15, MRPS16, MRPS17, MRPS21, MRPS23, MRPS27, MRPS28, DAP3, MRPS31, MRPS34, MRPS35, CHCHD1, AURKAIP1
Prostate adenocarcinomaUpregulationMRPL17
Thyroid cancerUpregulationMRPL14, MRPL17
DownregulationPTCD3
Uterine corpus endometrial carcinomaUpregulationMRPL3, MRPL12, MRPL13, MRPL14, MRPL15, MRPL17, MRPL19, MRPL47, MRPL51, MRPL52, GADD45GIP1, MRPS12, MRPS15, MRPS16, MRPS17, MRPS26, MRPS33, MRPS34, CHCHD1, AURKAIP1
DownregulationPTCD3
Uterine carcinosarcomaUpregulationMRPL3, MRPL11, MRPL12, MRPL13, MRPL14, MRPL15, MRPL17, MRPL47, MRPL51, MRPL53, MRPS6, MRPS12, MRPS15, MRPS16, MRPS34, CHCHD1, AURKAIP1
DownregulationPTCD3
* |log2 (FC)| > 1, q value < 0.01.
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

Wu, H.; Zhu, X.; Zhou, H.; Sha, M.; Ye, J.; Yu, H. Mitochondrial Ribosomal Proteins and Cancer. Medicina 2025, 61, 96. https://doi.org/10.3390/medicina61010096

AMA Style

Wu H, Zhu X, Zhou H, Sha M, Ye J, Yu H. Mitochondrial Ribosomal Proteins and Cancer. Medicina. 2025; 61(1):96. https://doi.org/10.3390/medicina61010096

Chicago/Turabian Style

Wu, Huiyi, Xiaowei Zhu, Huilin Zhou, Min Sha, Jun Ye, and Hong Yu. 2025. "Mitochondrial Ribosomal Proteins and Cancer" Medicina 61, no. 1: 96. https://doi.org/10.3390/medicina61010096

APA Style

Wu, H., Zhu, X., Zhou, H., Sha, M., Ye, J., & Yu, H. (2025). Mitochondrial Ribosomal Proteins and Cancer. Medicina, 61(1), 96. https://doi.org/10.3390/medicina61010096

Article Metrics

Back to TopTop