Spandidos Publications Logo
  • About
    • About Spandidos
    • Aims and Scopes
    • Abstracting and Indexing
    • Editorial Policies
    • Reprints and Permissions
    • Job Opportunities
    • Terms and Conditions
    • Contact
  • Journals
    • All Journals
    • Oncology Letters
      • Oncology Letters
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Oncology
      • International Journal of Oncology
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Molecular and Clinical Oncology
      • Molecular and Clinical Oncology
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Experimental and Therapeutic Medicine
      • Experimental and Therapeutic Medicine
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Molecular Medicine
      • International Journal of Molecular Medicine
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Biomedical Reports
      • Biomedical Reports
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Oncology Reports
      • Oncology Reports
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Molecular Medicine Reports
      • Molecular Medicine Reports
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • World Academy of Sciences Journal
      • World Academy of Sciences Journal
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Functional Nutrition
      • International Journal of Functional Nutrition
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Epigenetics
      • International Journal of Epigenetics
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Medicine International
      • Medicine International
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
  • Articles
  • Information
    • Information for Authors
    • Information for Reviewers
    • Information for Librarians
    • Information for Advertisers
    • Conferences
  • Language Editing
Spandidos Publications Logo
  • About
    • About Spandidos
    • Aims and Scopes
    • Abstracting and Indexing
    • Editorial Policies
    • Reprints and Permissions
    • Job Opportunities
    • Terms and Conditions
    • Contact
  • Journals
    • All Journals
    • Biomedical Reports
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Experimental and Therapeutic Medicine
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Epigenetics
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Functional Nutrition
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Molecular Medicine
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • International Journal of Oncology
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Medicine International
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Molecular and Clinical Oncology
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Molecular Medicine Reports
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Oncology Letters
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • Oncology Reports
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
    • World Academy of Sciences Journal
      • Information for Authors
      • Editorial Policies
      • Editorial Board
      • Aims and Scope
      • Abstracting and Indexing
      • Bibliographic Information
      • Archive
  • Articles
  • Information
    • For Authors
    • For Reviewers
    • For Librarians
    • For Advertisers
    • Conferences
  • Language Editing
Login Register Submit
  • This site uses cookies
  • You can change your cookie settings at any time by following the instructions in our Cookie Policy. To find out more, you may read our Privacy Policy.

    I agree
Search articles by DOI, keyword, author or affiliation
Search
Advanced Search
presentation
International Journal of Molecular Medicine
Join Editorial Board Propose a Special Issue
Print ISSN: 1107-3756 Online ISSN: 1791-244X
Journal Cover
December-2025 Volume 56 Issue 6

Full Size Image

Sign up for eToc alerts
Recommend to Library

Journals

International Journal of Molecular Medicine

International Journal of Molecular Medicine

International Journal of Molecular Medicine is an international journal devoted to molecular mechanisms of human disease.

International Journal of Oncology

International Journal of Oncology

International Journal of Oncology is an international journal devoted to oncology research and cancer treatment.

Molecular Medicine Reports

Molecular Medicine Reports

Covers molecular medicine topics such as pharmacology, pathology, genetics, neuroscience, infectious diseases, molecular cardiology, and molecular surgery.

Oncology Reports

Oncology Reports

Oncology Reports is an international journal devoted to fundamental and applied research in Oncology.

Experimental and Therapeutic Medicine

Experimental and Therapeutic Medicine

Experimental and Therapeutic Medicine is an international journal devoted to laboratory and clinical medicine.

Oncology Letters

Oncology Letters

Oncology Letters is an international journal devoted to Experimental and Clinical Oncology.

Biomedical Reports

Biomedical Reports

Explores a wide range of biological and medical fields, including pharmacology, genetics, microbiology, neuroscience, and molecular cardiology.

Molecular and Clinical Oncology

Molecular and Clinical Oncology

International journal addressing all aspects of oncology research, from tumorigenesis and oncogenes to chemotherapy and metastasis.

World Academy of Sciences Journal

World Academy of Sciences Journal

Multidisciplinary open-access journal spanning biochemistry, genetics, neuroscience, environmental health, and synthetic biology.

International Journal of Functional Nutrition

International Journal of Functional Nutrition

Open-access journal combining biochemistry, pharmacology, immunology, and genetics to advance health through functional nutrition.

International Journal of Epigenetics

International Journal of Epigenetics

Publishes open-access research on using epigenetics to advance understanding and treatment of human disease.

Medicine International

Medicine International

An International Open Access Journal Devoted to General Medicine.

Journal Cover
December-2025 Volume 56 Issue 6

Full Size Image

Sign up for eToc alerts
Recommend to Library

  • Article
  • Citations
    • Cite This Article
    • Download Citation
    • Create Citation Alert
    • Remove Citation Alert
    • Cited By
  • Similar Articles
    • Related Articles (in Spandidos Publications)
    • Similar Articles (Google Scholar)
    • Similar Articles (PubMed)
  • Download PDF
  • Download XML
  • View XML
Review Open Access

Research progress on TMEM proteins in cancer progression and chemoresistance (Review)

  • Authors:
    • Ji Shi
    • Duo Zheng
    • Bing Yao
    • Qiang Liu
    • Huizhe Xu
    • Haozhe Piao
  • View Affiliations / Copyright

    Affiliations: Department of Neurosurgery, Liaoning Cancer Hospital and Institute, Shenyang, Liaoning 110042, P.R. China, Department of Neurology, Shengjing Hospital of China Medical University, Shenyang, Liaoning 110042, P.R. China, Department of Neurosurgery, Liaoning Cancer Hospital and Institute, Shenyang, Liaoning 110042, P.R. China, Central Laboratory Department, Liaoning Cancer Hospital and Institute/Cancer Hospital of Dalian University of Technology/Cancer Hospital of China Medical University, Shenyang, Liaoning 110042, P.R. China
    Copyright: © Shi et al. This is an open access article distributed under the terms of Creative Commons Attribution License.
  • Article Number: 219
    |
    Published online on: October 10, 2025
       https://doi.org/10.3892/ijmm.2025.5660
  • Expand metrics +
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Metrics: Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )
Cited By (CrossRef): 0 citations Loading Articles...

This article is mentioned in:



Abstract

In cancer research, the transmembrane (TMEM) family of proteins has attracted considerable attention due to its role in tumor progression and chemoresistance. These membrane proteins are integral to cellular processes, including signal transduction, ion transport and cellular homeostasis, rendering them promising therapeutic targets. The TMEM proteins are implicated in several types of cancer, including breast, ovarian, lung and thyroid cancer, where they regulate numerous cellular processes, including proliferation, migration, invasion and survival. Notably, TMEM45A and TMEM158 contribute to resistance to platinum‑based chemotherapy by increasing the expression of proteins associated with hypoxic conditions and multidrug resistance. Additionally, epigenetic regulation, particularly promoter methylation of TMEM88, is pivotal in regulating TMEM88 expression and function in chemoresistance. The present review presents a systematic and comprehensive overview of the structural features, biological functions and regulatory mechanisms of key TMEM proteins across various types of cancer. It also highlights emerging connections between TMEM proteins and the tumor microenvironment, emphasizing their potential as promising therapeutic targets. The novel findings underscore the key role of the TMEM protein family in overcoming chemoresistance and lay a foundation for the development of targeted therapeutic strategies in cancer treatment.

Introduction

Membrane proteins (MPs) are proteins integrated into biological membranes, where they interact with lipid domains to execute a range of essential functions within living organisms (1). These proteins are key for maintaining the structural integrity and functional operation of cells, engaging in numerous processes such as ion transport, signal transduction and regulation of cellular homeostasis (2). MPs are central to the regulation of ion gradients across membranes, modulation of cellular communication and responsiveness to extracellular signals (3). In addition to their physiological roles, MPs also carry out a pivotal role in the development of multidrug resistance (MDR), rendering them central to mechanisms of therapeutic resistance (2). Research has expanded the understanding of MPs, revealing that their functions extend beyond basic cellular roles to include key involvement in complex physiological processes and disease states, particularly in different types of cancer such as colorectal, ovarian, lung breast (1-5).

In the majority of mammalian cells, MPs assist in maintaining membrane structure, facilitate the transport of ions, nutrients and waste products or facilitate intercellular communication (3). They are integral to the formation of cellular microenvironments that enable efficient communication between cells, regulating processes such as immune responses, tissue remodeling and apoptosis (6). MPs are also involved in the establishment of the blood-brain barrier, the regulation of neurotransmitter transport and cellular metabolic processes (7). This array of functions underscores their key roles in maintaining cellular homeostasis and regulating the essential biological processes necessary for life.

Emerging research has emphasized the key role of MPs in the pathogenesis of cancer, including breast cancer, colorectal cancer, lung cancer, ovarian cancer and brain tumors (1,3-7). For example, subfamilies of tyrosine kinase receptors, including epidermal growth factor receptors (EGFRs) and fibroblast growth factor receptors (FGFRs), are key in signaling pathways that regulate cell proliferation, survival and migration, all important processes that drive tumorigenesis (8,9). These receptors become activated upon binding of ligands to their respective receptors, inducing conformational changes that initiate signal transduction to the cytoplasm via processes such as dimerization (10). Upon activation, receptors trigger downstream signaling cascades that regulate gene expression, cell cycle progression and metabolic reprogramming (11). In cancer cells, dysregulation of receptor signaling contributes to tumor growth, metastasis and resistance to chemotherapy, highlighting MPs as potential therapeutic targets for the management of cancer, including breast cancer and pancreatic cancer (12,13). MPs carry out a key role not only in maintaining normal cellular activities but also in cancer development and chemotherapy resistance (3-5). The tyrosine kinase receptor family, including EGFR and FGFR, has garnered increasing attention for its role in regulating cancer cell proliferation, survival and migration, particularly in the context of chemotherapy resistance (3-5,8,9). Cancer cells often exploit dysregulated membrane protein signaling pathways to drive tumor growth and metastasis, with resistance mechanisms frequently associated with alterations in membrane protein function (13-15). For example, discoidin domain receptor 2 interactions with collagen type I promote breast cancer cell proliferation and enhance resistance to chemotherapy, highlighting its potential role in chemoresistance in breast cancer (13-15). Moreover, transmembrane proteins (TPs) such as transmembrane (TMEM) 45A, TMEM158, TMEM88, TMEM205 and TMEM16A have been identified as key mediators of chemoresistance in several types of cancer (4,5,13,14). These proteins regulate cell proliferation, inhibit apoptosis or modulate the tumor microenvironment (TME) to increase resistance of cancer cells to chemotherapeutic agents (1,4-6,13,14). Therefore, MPs not only serve as key targets in cancer therapy but can also help in understanding and addressing chemotherapy resistance mechanisms.

The complexity of the nexus of signal transduction mediated by MPs presents substantial challenges in elucidating their precise roles in cellular processes and disease mechanisms (16). One of the primary challenges in research on MPs lies in establishing the structure-function relationship that governs ligand activation and enzyme activity specificity (17). Structural studies of MPs have been hindered by their amphipathic nature, which complicates their isolation and analysis in vitro (1,3-5,18). However, recent advances in cryo-electron microscopy (cryo-EM) and other high-resolution imaging techniques have enabled the structural elucidation of key MPs, offering insights into their functional mechanisms at the atomic level (19). These technological advances have markedly improved the ability to study MPs in their natural environments, thus facilitating the development of targeted drugs with high specificity and efficacy (19,20).

MPs are not only essential for cellular function but also act as key therapeutic targets, making them central to drug discovery efforts (21). They are involved in a range of physiological processes, many of which can be modulated by therapeutic agents targeting MPs to achieve clinical effects. Drug discovery efforts have concentrated on identifying small molecules, large molecules, peptides, polysaccharides and antibodies, particularly those that can specifically interact with MPs to regulate their activity (22). For example, monoclonal antibodies targeting EGFR and human epidermal growth factor receptor 2 (HER2) have been developed successfully and are now used in the treatment of several types of cancer, including glioma and breast cancer (22,23). Large molecule drugs, including polysaccharides and peptides, are seen as promising alternatives to traditional small molecule drugs, offering various therapeutic benefits such as reduced toxicity and enhanced efficacy when used in combination with conventional chemotherapy. For example, polysaccharides derived from medicinal plants such as Anemarrhena asphodeloides and Houttuynia cordata have demonstrated notable bioactivities, including immune modulation and anti-inflammatory effects, which can enhance the effectiveness of cancer treatments, including lung cancer and liver cancer (24,25). Furthermore, recent studies have highlighted the ability of peptides to target specific cancer cell pathways, enhancing treatment outcomes in several types of cancer, including pancreatic cancer (PC) (14,26). These findings underscore the potential of utilizing large-molecule drugs in cancer therapies, particularly in addressing the challenge of chemoresistance. Furthermore, several PMPs, including ion channels, transporters and G protein-coupled receptors (GPCRs), are well-established therapeutic targets for diseases ranging from cardiovascular disorders to neurological conditions (27-29). Given the key role of MPs in the pathophysiology of numerous diseases, understanding their structural and functional properties is essential for the development of novel drugs and therapies.

Biological processes, particularly those involving MPs, constitute a complex, interconnected network within the cell. Certain MPs can modulate interactions between various signaling pathways that govern fundamental cellular processes, such as metabolism, apoptosis and immune responses (30,31). The cyclic nature of these networks is particularly key for understanding how drugs influence cellular processes. The majority of therapeutic agents exert their effects by binding to MPs, thereby modulating signaling pathways essential for disease progression (32). Therefore, MPs serve as promising targets for novel therapeutic strategies targeting specific disease mechanisms at the molecular level. The expanding body of research on MPs and their interactions with drugs continues to reveal novel opportunities for drug development and personalized therapies (21,33).

MPs are essential to cellular function and disease pathogenesis (34). The study of MPs is essential for advancing the understanding of both normal physiology and disease mechanisms. Given their roles in signaling, cellular maintenance and disease progression, MPs serve as key targets for therapeutic intervention (7,12). Ongoing advancements in structural biology, coupled with novel drug discovery technologies, highlight novel opportunities for exploiting MPs as drug targets. Research on MPs holds potential for developing innovative treatments for a range of diseases, including cancer, cardiovascular disorders and neurological conditions (35-37).

Structure and function of the TMEM proteins

MPs are classified according to their interactions with membranes, which are categorized into peripheral MPs, integral MPs and lipid-anchored MPs (38). Integral MPs contain ≥1 transmembrane segment, commonly referred to as a TP (4). TMEM proteins represent a subgroup of integral MPs, defined by their ability to traverse the lipid bilayer, thus establishing stable anchors within the membrane. These proteins are essential in the transport of molecules across the cellular membrane (5,39). A protein is classified as a member of the TMEM family if it contains at least one predicted transmembrane domain that spans the biological membrane, either partially (in monomeric form) or fully (in multimeric form) (4).

Due to their amphipathic nature, members of the TMEM protein family adopt specific secondary structures when incorporation into the membrane (40,41). Based on their different structures, TMEMs can be divided into two categories: α-helical proteins and β-barrel proteins (Fig. 1). These two structures are the dominant structures of all transmembrane proteins, with α-helical proteins being mostly found in the cytoplasm and subcellular septum (Fig. 1A-C), while β-barrel proteins are mostly found in chloroplasts, bacteria and mitochondrial membranes (Fig. 1D) (40,41). These structural features allow TMEM proteins to modulate essential processes, including nutrient transport, waste removal and signal transduction (40,41).

Classification of transmembrane
protein structures. (A) Single-pass membrane protein, the lipid
bilayer is traversed by the polypeptide chain via a single α-helix.
(B) A number of transmembrane proteins are predominantly situated
on the cytoplasmic side of the lipid bilayer. (C) Multi-membrane
protein, the lipid bilayer is traversed by the polypeptide chain
via two or more α-helix. (D) Some transmembrane proteins form large
ion channels by transversing the plasma membrane with multiple
β-folded sheets.

Figure 1

Classification of transmembrane protein structures. (A) Single-pass membrane protein, the lipid bilayer is traversed by the polypeptide chain via a single α-helix. (B) A number of transmembrane proteins are predominantly situated on the cytoplasmic side of the lipid bilayer. (C) Multi-membrane protein, the lipid bilayer is traversed by the polypeptide chain via two or more α-helix. (D) Some transmembrane proteins form large ion channels by transversing the plasma membrane with multiple β-folded sheets.

TMEM proteins exhibit considerable functional diversity. These proteins are classified into distinct functional groups, including GPCRs, ion channels, transporters, carriers and other receptors (Fig. 2). GPCRs, one of the largest and most versatile classes of TMEM proteins, are involved in a range of signaling pathways that regulate essential physiological processes, including sensory perception, immune responses and cellular communication (42). Notably, TMEM proteins, including GPCRs (Fig. 2A), carry out key roles in cancer cell signaling, influencing tumor progression and metastasis (43). Ion channels (Fig. 2B) are TPs that form pores, allowing ions (such as Na+, K+, Ca2+ and Cl−) to passively flow across the membrane. This movement is vital for processes such as nerve signal transmission, muscle contraction and maintaining cellular ion balance (44). Transporter proteins (Fig. 2C) facilitate the movement of molecules across the lipid bilayer, typically against their concentration gradients. This process requires energy and carries out a key role in nutrient uptake, waste removal and the maintenance of cellular homeostasis (43,44). Dysfunction of these proteins is implicated in a variety of diseases, including neurological and cardiovascular disorders (43,44). TMEM proteins also function as transmembrane receptors (Fig. 2D), such as the well-known EGFR, which is involved in regulating cell growth, differentiation and survival. Dysregulation of these transmembrane receptors is frequently associated with the development and progression of various types of cancer, including breast, lung, ovarian, cervical, bladder, esophageal, brain, and head and neck cancer (45,46).

Classification of transmembrane
family functions. (A) GPCRs mediate cellular responses by
activating intracellular signaling pathways through G-proteins upon
ligand binding. (B) Ion channels allow the passive flow of ions
across the membrane, essential for processes such as nerve
transmission and muscle contraction. (C) Transporter proteins
actively move molecules across the membrane, often against
concentration gradients, to maintain cellular homeostasis. (D)
Receptors bind specific molecules to initiate cellular responses,
influencing processes such as immune function, cell growth and
metabolism. GPCR, G protein-coupled receptor.

Figure 2

Classification of transmembrane family functions. (A) GPCRs mediate cellular responses by activating intracellular signaling pathways through G-proteins upon ligand binding. (B) Ion channels allow the passive flow of ions across the membrane, essential for processes such as nerve transmission and muscle contraction. (C) Transporter proteins actively move molecules across the membrane, often against concentration gradients, to maintain cellular homeostasis. (D) Receptors bind specific molecules to initiate cellular responses, influencing processes such as immune function, cell growth and metabolism. GPCR, G protein-coupled receptor.

Previous studies have underscored the role of TMEM proteins in various disease mechanisms, which highlights their potential as therapeutic targets for drug development (47,48). For example, mutations in TMEM proteins have been associated with several genetic disorders, including cystic fibrosis, leading to defects in ion channels and disruption of chloride transport (47,48). Moreover, TMEM proteins, such as glucose transporters (GLUTs), carry out a pivotal role in cancer metabolism by facilitating glucose uptake, therefore promoting tumor growth, especially under nutrient-limited conditions (49). This highlights the potential of targeting TMEM proteins for therapeutic interventions in a range of diseases.

The increasing understanding of the structural and functional properties of TMEM proteins offers potential for drug development. Their involvement in cellular signaling and molecular transport makes them strong candidates for targeted therapies (50). Advances in structural biology, particularly the application of cryo-EM, have facilitated the determination of the atomic structures of key TMEM proteins, offering valuable insights into their functional mechanisms (51-53). The TMEM family proteins are essential components of the cellular membrane system, carrying out key roles in cellular signaling and the transport of ions and small molecules (39). Specific TMEM proteins interact with hormone or neurotransmitter receptors, inducing structural changes that initiate signaling cascades important for cellular function. These proteins mediate the selective transfer of ions or molecules across biological membranes, thereby establishing concentration gradients essential for cellular homeostasis and function (54). TMEM proteins are essential not only to the plasma membrane but also to the membranes of various intracellular organelles, including mitochondria, the endoplasmic reticulum (ER), lysosomes and the Golgi apparatus, where they participate in key biological processes (55). These processes carry out a key role in an array of physiological functions and diseases, including immune response regulation and tumorigenesis. Examples of key TMEM proteins include TMEM45A in epidermal keratinization (56), TMEM16 in smooth muscle contraction (57), TMEM165 in protein glycosylation (58) and TMEM9B in immune response modulation (59).

TMEM proteins are key contributors to cancer biology, carrying out key roles in tumor initiation, progression and metastasis. Their differential expression in various types of cancer not only serves as biomarkers but also offers new targets for therapeutic intervention (4,60). Furthermore, alterations in TMEM protein expression levels or function are closely associated with tumor progression, metastasis and drug resistance (60). Recent studies have emphasized the vital role of TMEM proteins in cancer cell survival and proliferation under harsh conditions, such as hypoxia or nutrient deprivation, which are commonly associated with the TME (4,5,12-16,60,61). For example, TMEM16, an ion channel, has been shown to regulate chloride and calcium ion flux, thereby influencing cell proliferation and migration in cancer cells (62). Likewise, TMEM65 has been implicated in regulating autophagy, an essential process for maintaining cellular homeostasis and promoting cancer cell survival (63). Given their upregulated expression in various types of cancer, targeting specific TMEM proteins presents a viable strategy for therapeutic intervention. Therapeutic strategies may include the development of small-molecule inhibitors or monoclonal antibodies that selectively target these proteins, thereby inhibiting their function and disrupting the signaling pathways they regulate (64). These approaches may address certain limitations of conventional cancer therapies, such as drug resistance and off-target effects, offering a more personalized and effective treatment regimen (65).

The TMEM protein family is important for maintaining cellular homeostasis and regulating various physiological processes. Their involvement in the pathogenesis of cancer and other diseases makes them potential candidates for targeted therapies. Ongoing research into the molecular mechanisms of TMEM proteins and their roles in disease progression is likely to lead to novel therapeutic strategies, particularly for several types of cancer and other disorders associated with abnormal TMEM protein activity.

Research on the correlation between TMEM proteins and cancer

TMEM family members as tumor suppressor genes
TMEM7

Multiple regions on chromosome 3p are commonly affected by loss of heterozygosity in several types of cancer (66). TMEM7 is a candidate tumor suppressor gene located at the 3p21.3 region of the human genome. It is a 232-amino acid, single-pass membrane protein, primarily expressed in the liver (67) and is integral to liver carcinogenesis (Fig. 3). It shares considerable sequence homology with the 28 kDa interferon-α (IFN-α)-responsive protein, suggesting a potential role in immune response modulation (67). TMEM7 has been implicated as a candidate tumor suppressor gene due to its frequent downregulation or inactivation in several types of cancer. Zhou et al (67) examined primary hepatocellular carcinoma (HCC) tissues and HCC cell lines, and found that TMEM7 mRNA expression was downregulated or silenced in 85% of primary HCC samples and 33% of HCC cell lines. This recurrent loss of function suggests that TMEM7 carries out a key role in suppressing hepatocarcinogenesis (67). Mechanistically, the downregulation of TMEM7 is primarily driven by epigenetic modifications such as DNA methylation and histone deacetylation, rather than by genomic deletions or mutations (67-69). To the best of our knowledge, no homozygous deletions or mutations of TMEM7 have been identified in HCC tissues or cell lines.

List of TMEM protein structures.
Representative schematic diagrams of 16 TMEM proteins, TMEM7,
TMEM25, TMEM16A, TMEM45A, TMEM45B, TMEM48, TMEM65, TMEM88,
TMEM106A, TMEM140, TMEM158, TMEM159, TMEM173, TMEM176A, TMEM176B
and TMEM205, illustrating their predicted transmembrane structures,
subcellular localization and classification as single-pass or
multi-pass membrane proteins. TMEM, transmembrane.

Figure 3

List of TMEM protein structures. Representative schematic diagrams of 16 TMEM proteins, TMEM7, TMEM25, TMEM16A, TMEM45A, TMEM45B, TMEM48, TMEM65, TMEM88, TMEM106A, TMEM140, TMEM158, TMEM159, TMEM173, TMEM176A, TMEM176B and TMEM205, illustrating their predicted transmembrane structures, subcellular localization and classification as single-pass or multi-pass membrane proteins. TMEM, transmembrane.

Zhou et al (67) demonstrated that treatment with DNA methylation inhibitors (such as 5-aza-2'-deoxycytidine) and histone deacetylase inhibitors (such as trichostatin A) can restore TMEM7 expression in knockdown cell lines, confirming the role of epigenetic alterations in its silencing. Through the ectopic expression of TMEM7 in two HCC cell lines with TMEM7 knockdown, cell proliferation, colony formation and cell migration were inhibited, and tumor formation in nude mice was reduced. After 7 days of IFN-α treatment in two highly invasive HCC cell lines, TMEM7 expression was markedly increased and cell migration was inhibited, further establishing the role of TMEM7 in liver cancer metastasis (67). These findings also suggest that modulating TMEM7 expression via IFN-α may have therapeutic potential for certain patients with HCC. While the role of TMEM7 in HCC is well-documented, its interaction with other genes on chromosome 3p further expands its roles. In a study by Kholodnyuk et al (68) involving human chromosome 3-mouse fibrosarcoma hybrids, the transcription of TMEM7 and its neighboring genes (such as LTF and SLC38A3) was found to be impaired in renal cell carcinoma (RCC) cell lines, suggesting a broader tumor-suppressive role across multiple cancer types (68).

The epigenetic silencing of TMEM7 during hepatocarcinogenesis highlights the functional loss of TMEM7 as a key factor in liver cancer development, while its restoration or upregulation (by IFN-α, for example) offers potential therapeutic benefits (67-69). Understanding the precise role of TMEM7 in liver cancer progression may pave the way for targeted therapies aimed at improving outcomes for patients with this aggressive malignancy.

TMEM25

TMEM25, a member of the immunoglobulin superfamily, carries out a role in various cellular processes, including immune responses, growth factor signaling and cell adhesion (70,71). Its expression is markedly altered in several types of cancer, particularly in triple-negative breast cancer (TNBC), colorectal cancer (CRC) and clear cell RCC (ccRCC), making it a promising biomarker for diagnosis, prognosis and therapeutic targeting. The functional importance of TMEM25 in cancer is underscored by a study demonstrating its epigenetic regulation, particularly through DNA methylation and histone modifications, which affect its expression and role in tumorigenesis (71).

In colon adenocarcinoma, Hrašovec et al (70) observed a marked downregulation of TMEM25 mRNA in 68% of tumor tissues from patients with primary colon cancer. This downregulation was linked to hypermethylation of a specific CpG site within the 5' untranslated region of the TMEM25 gene, suggesting that epigenetic silencing through DNA methylation is a key mechanism responsible for the decreased expression of TMEM25 in colon adenocarcinoma. The inverse relationship between TMEM25 hypermethylation and its expression levels underscores the key role of epigenetic regulation in modulating TMEM25 expression during colon carcinogenesis (70). In breast cancer, the expression of TMEM25 is reduced in 50% of tumor samples compared with normal tissues, as reported by Doolan et al (72). Increased TMEM25 expression levels were associated with improved clinical outcomes, particularly in patients undergoing adjuvant chemotherapy, where it was associated with prolonged survival (72). By contrast, TMEM25 expression was often undetectable in TNBC, a subtype of breast cancer that is difficult to treat and associated with poor prognosis (70-72). In TNBC, TMEM25 acts as a suppressor of tumor progression by inhibiting the activation of the STAT3 signaling pathway. Specifically, TMEM25 physically interacts with the EGFR monomer, preventing EGFR-mediated STAT3 phosphorylation. This interaction sequesters unphosphorylated STAT3 in the cytoplasm, thereby inhibiting its nuclear translocation and subsequent tumor-promoting gene transcription (12,71,72). However, in TMEM25-deficient TNBC cells, EGFR monomers can phosphorylate STAT3 independently of ligand binding, leading to basal STAT3 activation and tumor progression. Restoring TMEM25 expression through adeno-associated virus delivery has been revealed to suppress STAT3 activation and TNBC progression in preclinical models, highlighting its potential as a therapeutic target (12,71,72). These findings indicate that TMEM25 may act as a valuable prognostic marker in breast cancer, particularly in predicting patient response to chemotherapy and providing insights into the molecular mechanisms underlying chemotherapy resistance. Growing evidence also indicates that TMEM25 carries out a tumor-suppressive role in ccRCC, a subtype of kidney cancer with limited therapeutic options (73). In ccRCC tissues, elevated DNA methylation of TMEM25 has been observed, suggesting epigenetic silencing in this type of cancer. Moreover, mutations in TMEM25 have been revealed to notably impact patient prognosis, with reduced TMEM25 expression associating with a more aggressive disease phenotype. Notably, the expression of TMEM25 in ccRCC is strongly associated with several immune-related factors, including immune checkpoint molecules, immune suppressive markers and major histocompatibility complex molecules, implying that TMEM25 may be involved in modulating the immune landscape of ccRCC (73). This interaction between TMEM25 and immune-related factors further underscores its role in immune evasion within the TME.

TMEM25 has been identified as a promising prognostic biomarker in colon cancer, breast cancer and ccRCC (70-72). Its expression is regulated epigenetically, primarily through DNA methylation and histone modifications, which suggests that reversing its epigenetic silencing could offer a promising therapeutic approach. Furthermore, the involvement of TMEM25 in immune modulation, especially through its interactions with immune checkpoint molecules, presents considerable promise for immunotherapeutic interventions. This establishes TMEM25 as a key therapeutic target in several types of cancer where its expression is reduced or absent. Given its pivotal role in cancer progression and immune regulation, TMEM25 may serve as a promising therapeutic target. Future research should focus on elucidating the molecular pathways through which TMEM25 influences tumor progression and immune responses. Restoring or enhancing TMEM25 expression could considerably improve treatment outcomes in colon cancer, breast cancer, and ccRCC, thereby positioning it as a promising target in future cancer therapies (70-72).

TMEM88

TMEM88 is a dual-transmembrane protein that participates in several key biological processes, particularly the regulation of human stem cell differentiation and embryonic development (74,75). Previous studies have indicated that TMEM88 carries out a key role in tumorigenesis, with its expression being altered in several types of cancer (74,75). Notably, TMEM88 binds to Dishevelled (DVL), preventing its interaction with LRP5/6 and inhibiting the canonical Wnt pathway. This leads to decreased β-catenin activity and downregulation of Wnt target genes, including c-Myc and cyclin D1 (75). While TMEM88 is widely expressed in various types of tumors, it is typically downregulated in the majority of tumor cells, suggesting its role as a potential tumor suppressor (76).

The expression of TMEM88 is also markedly decreased in thyroid cancer (TC) tissues and cell lines (75-77). Studies have revealed that TMEM88 inhibits the proliferation, colony formation and invasion of TC cancer cells, primarily through binding to DVL (75-77). In both in vitro and in vivo models, overexpression of TMEM88 has been shown to suppress tumorigenic properties, further supporting the role of TMEM88 as a tumor suppressor in TC. These findings suggest that restoring TMEM88 expression or inhibiting its methylation may offer a promising therapeutic strategy for treating TC. The involvement of TMEM88 in bladder cancer has been further emphasized instudies (78,79). Upregulation of TMEM88 in bladder cancer leads to a marked reduction in cellular proliferation and invasion (78,79). Bioinformatic analysis revealed that TMEM88 expression was considerably reduced in bladder cancer tissues compared with normal tissues (79). Zhao et al (79) provided evidence that loss of TMEM88 expression in bladder cancer cells enhances invasive and proliferative capacities, while re-expression of TMEM88 effectively reverses these phenotypes (79). Additionally, overexpression of TMEM88 in bladder cancer xenograft models inhibited tumor growth (79), indicating that TMEM88 acts as a tumor suppressor gene in bladder cancer as well.

The biological role of TMEM88 depends on its subcellular localization. Membrane-localized TMEM88 inhibits Wnt signaling, while cytosolic TMEM88 promotes tumor progression through alternative pathways. Zhang et al (80) reported that cytosolic localization of TMEM88 in non-small cell lung cancer (NSCLC) is associated with poor differentiation, a high TNM stage, lymph node metastasis and worse overall survival (OS). TMEM88 was shown to promote invasion and metastasis by activating the p38-GSK3 β-Snail signaling pathway. Overexpression of TMEM88 in NSCLC cells inhibited excessive cell proliferation, invasion and migration, leading to the prevention of tumor growth in xenograft models (80). Furthermore, hypermethylation of the TMEM88 promoter has been observed in NSCLC tissues, which is associated with a worse prognosis. Demethylation of TMEM88 resulted in the inhibition of cell proliferation, migration and invasion, suggesting that TMEM88 expression and its epigenetic regulation carry out a key role in NSCLC progression (81). In TNBC, TMEM88 overexpression in the cytosol stimulated invasion and metastasis by interacting with DVL proteins and promoting Snail expression while inhibiting tight junction proteins, such as zonula occludens-1 and occluding (82). Thus, increased TMEM88 expression was associated with improved prognosis in HCC and bladder cancer, while its cytosolic localization was predictive of a worse outcome in NSCLC and TNBC.

TMEM88 carries out a key role in regulating cell proliferation, migration and invasion, and has demonstrated notable potential as both a diagnostic biomarker and therapeutic target across various types of cancer (79-82). Its expression is frequently downregulated in tumors, and promoter hypermethylation of TMEM88 is associated with a worse prognosis in NSCLC (80,81). These findings highlight TMEM88 as a potential marker for disease progression and survival prediction. The epigenetic regulation of TMEM88, particularly through DNA methylation, offers a novel therapeutic avenue. Demethylating agents can restore TMEM88 expression (81), thereby suppressing tumor cell growth and metastasis. As a negative regulator of the Wnt/β-catenin signaling pathway, TMEM88 exhibits tumor-suppressive activity in NSCLC, TC and bladder cancer. Therapeutic strategies aimed at restoring or enhancing TMEM88 expression could thus represent a viable approach for cancer treatment. Ongoing research into the molecular mechanisms of TMEM88 is expected to further clarify its value as both a therapeutic target and a prognostic biomarker in clinical oncology.

TMEM106A

TMEM106A is located on chromosome 17 and consists of 262 amino acids. TMEM106A is a conserved type II TP that has been implicated in suppressing tumorigenesis across multiple types of cancer (83,84). TMEM106A is typically expressed in normal gastric tissue and is often downregulated or silenced in gastric cancer (GC), due to methylation of the promoter region, and its methylation is associated with smoking and tumor metastasis (83,84). Frequent methylation of TMEM106A in GC tissues highlights its potential role as a tumor suppressor. Xu et al (84) demonstrated that overexpression of TMEM106A markedly inhibited GC cell proliferation and induced apoptosis, while also delaying tumor growth in xenograft models (84). Mechanistic studies suggested that these effects were closely associated with the activation of caspase-2, caspase-9 and caspase-3, as well as the inactivation of poly(ADP-ribose) polymerase (PARP) (84). Thus, TMEM106A has emerged as a functional tumor suppressor in the pathogenesis of GC.

Methylation of TMEM106A is generally associated with its reduced expression and the promoter region methylation serves as a key regulatory mechanism in GC. The downregulation of TMEM106A expression in patients with GC is associated with a worse prognosis, emphasizing its clinical relevance as a prognostic biomarker (85). In renal cancer, TMEM106A expression is markedly reduced, and its restoration in cancer cells leads to attenuated proliferation, reduced migration and enhanced caspase-3-dependent apoptosis (86). Similarly, the expression of TMEM106A in HCC tissues is notably reduced compared with normal liver tissue, accompanied by frequent promoter methylation in HCC tissues, further reinforcing its potential as a tumor suppressor (87). Treatment with 5-Aza-2'-deoxycytidine in HCC cells with high methylation restored TMEM106A expression and markedly suppressed the malignant behaviors of these cells, including reduced cell migration, invasion and lung metastasis (87). In vivo experiments demonstrated that overexpression of TMEM106A in HCC cells resulted in considerable tumor suppression, and this was associated with the suppression of epithelial-mesenchymal transition (EMT) and the inactivation of the ERK1/2/Slug signaling pathway (87). Moreover, in NSCLC, TMEM106A expression was revealed to be markedly decreased in tumor tissues. Overexpression of TMEM106A in NSCLC cells reduced proliferation, migration and invasion, while inducing apoptosis. It also repressed EMT by modulating E-cadherin, N-cadherin and vimentin expression, and inhibited the PI3K/Akt/NF-κB signaling pathway (88). In urothelial carcinoma, TMEM106A was identified as a novel biomarker, with its methylation levels in urine samples revealing high diagnostic accuracy for early detection (89).

TMEM106A also exhibits antiviral activity, specifically in the IFN-mediated antiviral immune response. TMEM106A was revealed to interact with the receptor SCARB2 of EV-A71 and CV-A16, thus disrupting viral binding to host cells and preventing viral infection (90). This finding not only highlights a novel mechanism of TMEM106A in antiviral immunity but also proposes a potential therapeutic target for antiviral strategies. In addition to its antiviral function, TMEM106A regulates macrophage activation and immune responses, especially in bacterial infection-induced inflammation. Its downregulation enhanced M1 macrophage polarization and exacerbated the LPS-induced inflammatory response through the MAPK and NF-κB signaling pathways, underscoring the key role of TMEM106A in immune homeostasis (91).

TMEM106A carries out a key tumor-suppressive role in various types of cancer, such as gastrointestinal tumors, liver cancer and NSCLC, with its dysregulated expression, including methylation, potentially serving as a valuable clinical prognostic biomarker (84,86-91). Furthermore, the diverse functions of TMEM106A in immune responses and antiviral defense establish it as a potential target for future cancer treatments and immune modulation strategies. Further research should prioritize elucidating the specific mechanisms by which TMEM106A operates in various types of cancer and exploring its clinical applications.

TMEM173

TMEM173, also known as stimulator of interferon genes (STING), is a key TP located on the outer membrane of the ER and carries out a key role in cancer immunology and immunotherapy. It is essential in innate immunity and is primarily expressed in hematopoietic cells and immune organs such as the thymus and spleen (92,93). Composed of four transmembrane domains, TMEM173 forms dimers that allow ligand binding and activation of immune responses, in addition to mediating reactions to cytosolic DNA, which can result from infections, cellular stress or genomic instability (94). TMEM173 carries out a key role in host defense against pathogenic infections by regulating type I IFN (IFN-I) signaling, pro-inflammatory cytokines and innate immunity (95). In addition to its role in innate immunity, TMEM173 also contributes to autoimmune inflammation, oxidative stress and autophagy (96).

While TMEM173 has demonstrated tumor-suppressive functions in various cancer types, including CRC, prostate cancer and melanoma (92); it is important to note that TMEM173 activation can act as a double-edged sword in cancer. It promotes antitumor immunity by enhancing dendritic cell (DC) maturation, T cell activation and macrophage polarization toward an M1 phenotype, which supports tumor elimination (97-99). T cell infiltration into tumors is largely dependent on IFN-I signaling, and activation of TMEM173 enhances immune surveillance against tumors. For example, TMEM173 agonists have been revealed to amplify the cGAS-STING pathway, leading to increased IFN-β secretion and T cell-mediated immune responses in colorectal and breast cancer models (97). Additionally, TMEM173 activation can reprogram tumor-associated macrophages (TAMs) from an immunosuppressive M2 phenotype to an immunostimulatory M1 phenotype, further enhancing antitumor immunity (97,100). TMEM173 expression is downregulated in various types of cancer, including GC and HCC, and this downregulation associates with tumor size, invasiveness and lymph node metastasis. In vitro experiments have revealed that knockout of TMEM173 promotes clonal formation, migration and invasion of GC cells, whereas its activation induces apoptosis and autophagy in malignant cells, inhibiting tumor growth (101). However, STING signaling can also promote immune evasion and tumor progression. For example, TMEM173 activation in tumor monocytes induces programmed death (PD)-ligand 1 expression, which suppresses T cell activity and contributes to resistance against TMEM173 agonist therapy (102). Similarly, TMEM173 signaling can expand regulatory B cells that produce IL-35, which attenuates natural killer (NK) cell function and compromises antitumor immunity (103). TMEM173 may carry out a key role in metastasis formation. Activation of TMEM173 promotes metastatic tumor formation and spread, partly through the production of TNF-α and other cytokines (104). For example, Bu et al (105) reported that, in clinical samples from patients with hepatitis B virus (HBV)-induced liver cancer, TMEM173 downregulation may contribute to immune evasion and subsequently facilitate tumor progression.

TMEM173 is not only a key regulator of tumor immunity but also holds substantial therapeutic potential within cancer immunotherapy. Targeting TMEM173, either by modulating its upstream or downstream signaling pathways, may provide novel therapeutic strategies to enhance cancer treatment outcomes, particularly when combined with immune checkpoint blockade therapies (106). Therefore, TMEM173 agonists are emerging as promising candidates for cancer immunotherapy. Nanoparticle-based delivery systems have been developed to enhance the spatiotemporal activation of TMEM173, improving its therapeutic efficacy while minimizing off-target effects (107). For example, TME-responsive nanoparticles have been designed to simultaneously activate TMEM173 and TLR4, leading to enhanced IFN-β production and T cell responses in colorectal and metastatic breast cancer models (97). Additionally, combining TMEM173 agonists with immune checkpoint inhibitors, such as anti-PD-1 antibodies, has shown synergistic effects in preclinical models (97,100). However, challenges remain in translating TMEM173 agonists to clinical practice. Issues such as poor drug stability, limited tumor targeting and immune-related adverse events need to be addressed (103,108). Nanomedicine approaches, including the use of radiosensitizers and targeted delivery systems, are being explored to overcome these limitations.

TMEM173 is a central player in cancer immunology, with its activation driving both antitumor immunity and immune evasion. While TMEM173 agonists hold potential for cancer immunotherapy, their clinical translation requires careful consideration of the TME and potential off-target effects. Future research should focus on developing targeted delivery systems and combination therapies to maximize the therapeutic potential of TMEM173 activation in cancer treatment.

TMEM176A/TMEM176B

TMEM176A and TMEM176B, both belonging to the membrane-spanning 4-domains family, have emerged as key regulators of immune responses and cancer progression. TMEM176A is localized on chromosome 7q36.1 and has been identified as a potential tumor suppressor in various types of cancer, including esophageal squamous cell carcinoma (ESCC) and CRC (109). In addition, TMEM176B, also known as tolerance-related and induced, carries out a key role in immune modulation and cancer progression (110).

TMEM176A has been revealed to suppress the growth of ESCC cells both in vitro and in vivo (110-112), suggesting its potential as a diagnostic and prognostic biomarker for ESCC. In ESCC, the loss of TMEM176A expression is associated with promoter hypermethylation (111). TMEM176A promoter methylation occurs in ~66% of primary ESCC tumors, which associates with its reduced expression, poor tumor differentiation, reduced survival rates and aggressive cancer phenotypes (111). The restoration of TMEM176A expression using demethylation agents, such as 5'-aza-2'-deoxycytidine, inhibits cell migration and invasion, thereby promoting apoptosis in ESCC cells and suppressing cancer progression. Additionally, in CRC, TMEM176A promoter methylation is observed in nearly half of the primary tumors, with TMEM176A overexpression resulting in suppressed cell proliferation, migration, invasion and induced apoptosis (112). Notably, TMEM176A is also involved in regulating glioma cell cycles through pathways such as ERK1/2, further validating its role as a tumor suppressor (112). Collectively, these findings establish TMEM176A as a potential tumor suppressor gene, with its expression levels serving as valuable prognostic indicators across multiple cancer types (113).

TMEM176B has been extensively studied in various types of cancer, where it often promotes tumor progression. In CRC, TMEM176B inhibits the NOD-, LRR-, and pyrin domain-containing protein 3 inflammasome, thereby suppressing NK cell activity and promoting EMT, a key process in metastasis. Silencing TMEM176B reduces tumor growth, metastasis and EMT while enhancing apoptosis and immune activation (114). In GC, TMEM176B overexpression drives tumor progression by regulating the PI3K-Akt-mTOR signaling pathway and amino acid metabolism. Its knockdown inhibits cell proliferation, migration and invasion, while promoting apoptosis. Clinically, TMEM176B expression is associated with a worse prognosis, making it a potential prognostic marker (115). TMEM176B also carries out a role in ovarian cancer (OC), where it acts as a tumor suppressor. Its downregulation is associated with increased EMT, proliferation and metastasis, mediated through the Wnt/β-catenin signaling pathway. Restoring TMEM176B expression inhibits these processes, highlighting its therapeutic potential (116).

TMEM176B also carries out a key role in regulating immune responses, especially in DCs (117). TMEM176B is recognized as a key modulator of allograft tolerance and immune system function (117). A study involving autologous tolerogenic DCs (ATDCs) demonstrated that TMEM176B expression in ATDCs was essential for inducing donor-specific immune tolerance and prolonging graft survival (118). TMEM176B-deficient ATDCs failed to cross-present antigens effectively, impairing their ability to induce regulatory CD8+ T cells and hindering allograft tolerance (118). TMEM176B inhibits inflammasome activation, which dampens anti-tumor immunity. Targeting TMEM176B unleashes inflammasome activity, increasing CD8+ T cell-mediated tumor control and enhancing the effects of anti-CTLA-4 and anti-PD-1 therapies (119). These findings underscore the importance of TMEM176B in modulating immune responses, particularly through its influence on antigen cross-presentation and regulation of phagosomal pH, both of which are important for immune tolerance (120). Additionally, TMEM176B has been implicated in regulating immune checkpoint responses, particularly in the context of PD-1 blockade therapy. Elevated TMEM176B expression in tumor stromal cells is associated with a worse prognosis, suggesting that it may act as a negative modulator of anti-tumor immunity (121). Its role in immune evasion has prompted investigations into its potential as a biomarker for predicting patient response to immune checkpoint inhibitors, including anti-PD-1 therapies. TMEM176B negatively regulated the antitumor activity of CD146+ TAMs. Inhibition of TMEM176B enhanced TAM-mediated anti-tumor immunity, highlighting a potential immunotherapeutic strategy (122). Thus, as an immune-regulating cation channel, TMEM176B modulates immune cell trafficking and function, further highlighting its potential as a therapeutic target (123).

TMEM176A and TMEM176B are known to physically interact and co-localize at both the plasma membrane and vesicular compartments within immune cells (124). In lymphoma, overexpression of both proteins has been associated with immune evasion, enabling cancer cells to evade immune surveillance and develop resistance to chemotherapy (124,125). This interaction suggested that the two proteins may function synergistically to regulate immune responses and contribute to tumor progression. Both TMEM176A and TMEM176B are downregulated during DC maturation, with their overexpression inhibiting this process and suggesting a role in maintaining an immature state that facilitates immune tolerance (123).

TMEM176A and TMEM176B have emerged as key factors in cancer biology, particularly in immune regulation and tumor suppression. Their roles in immune modulation, especially through the regulation of DC function and antigen cross-presentation, position them as promising targets for novel therapeutic strategies aimed at enhancing immune checkpoint blockade efficacy and improving cancer immunotherapy outcomes. Future research should focus on more comprehensively elucidating the precise molecular mechanisms by which TMEM176A and TMEM176B regulate immune responses and tumor progression. Clinical trials targeting these proteins, either alone or in combination with existing immune checkpoint inhibitors, may provide key insights into their therapeutic potential.

TMEM family members as oncogenes
TMEM16A

TMEM16 proteins, also known as anoctamins, are a family of ten MPs with various tissue expression and subcellular localization. TMEM16A, also referred to as anoctamin 1 (ANO1), is a calcium-activated chloride channel (CaCC) mapped to chromosome 11q13. It contains eight transmembrane domains and a highly conserved, functionally undefined domain (DUF590), with its expression predominantly observed in secretory epithelial cells, smooth muscle tissue and sensory neurons (126). TMEM16A is acknowledged as a key modulator of tumor progression. It is frequently amplified in a variety of malignancies, including breast cancer, esophageal cancer and head and neck squamous cell carcinomas (HNSCCs), where its expression is associated with a poor prognosis, increased tumor size and advanced clinical stages (127).

TMEM16A carries out a key role in tumorigenesis by activating key signaling pathways, including the EGFR and calcium/calmodulin-dependent protein kinase pathways (128). Silencing of TMEM16A has been revealed to inhibit the proliferation of esophageal and HNSCC cells and induce apoptosis through the MAPK and AKT signaling pathways. TMEM16A is involved in EMT, a key process in cancer metastasis, by regulating the transition from an epithelial to a mesenchymal state, thus enhancing cell migration and invasion (129). In CRC, TMEM16A carries out a key role in disease progression by promoting cell migration and invasion. It activates the Wnt/β-catenin signaling pathway by upregulating key components such as Frizzled protein 1 and β-catenin, while downregulating GSK3β (129,130). The overexpression of TMEM16A is notably elevated in HNSCC, esophageal and prostate cancer, where it is associated with distant metastasis and a poor prognosis (131).

The molecular mechanisms underlying these effects include a role independent of its channel function in promoting cell proliferation. Research suggests that TMEM16A regulates cancer cell growth by interacting with other MPs rather than exclusively through its chloride conductance (131). For example, TMEM16A has been revealed to interact with EGFR, thereby potentiating EGFR signaling in HNSCC cells (128). This interaction not only stabilizes EGFR but also increases TMEM16A protein expression, thereby establishing a feedback loop between TMEM16A and EGFR signaling pathways, which likely contributes to enhanced cancer cell proliferation and survival. Knockdown of TMEM16A has been revealed to reduce myeloid cell leukemia 1 expression and promote the perinuclear redistribution of p27Kip1, thereby inducing cell cycle arrest, suppressing tumor proliferation and promoting apoptosis (132). However, Shiwarski et al (133) reported that stable reduction of TMEM16A expression enhanced the motility of HNSCC cells and facilitated metastasis, independent of 11q13 amplification. Additionally, TMEM16A overexpression is associated with increased Erk1/2 activity, reduced Bim expression and decreased apoptotic activity, contributing to tumor progression and cisplatin resistance (134). These conflicting findings underscore the multifaceted role of TMEM16A in HNSCC progression. Further studies revealed that TMEM16A expression is associated with tumor subtype: In general, in HNSCC, elevated TMEM16A levels associate with shorter survival and increased risk of distant metastasis (135,136), whereas in human papilloma virus (HPV)-positive HNSCC, its expression was not associated with prognostic significance (136). This suggests that TMEM16A-targeted therapies may be inappropriate for HPV-positive patients. Collectively, these findings indicate that the function of TMEM16A is context-dependent, shaped by the TME and specific cellular subtypes. Continued clinical and molecular investigations are required to clarify its precise role in cancer progression.

TMEM16A has emerged as a potential biomarker for several types of cancer, including esophageal and gastrointestinal squamous cell carcinoma (127). Its expression is associated with clinical staging and disease progression, making it a valuable prognostic indicator. Moreover, TMEM16A functions as a predictor of therapeutic response to EGFR-targeted therapies, particularly in HNSCC, where TMEM16A overexpression may mediate resistance to EGFR inhibitors (137). Given its key role in tumor progression and interaction with EGFR signaling, TMEM16A represents a compelling therapeutic target. Targeting TMEM16A in conjunction with EGFR may offer a strategy to bypass resistance mechanisms and enhance the efficacy of existing EGFR-targeted therapies in cancer treatment (128).

Amplification of chromosome 11q13, where TMEM16A (ANO1) is located, is associated with a poor prognosis in patients with breast cancer. TMEM16A amplification and overexpression are associated with a higher tumor grade and worse outcomes, primarily through activation of EGFR and CaMKII signaling pathways to promote cell proliferation (138). TMEM16A expression also associates with immunohistochemical markers such as β-catenin, cyclin D1 and E-cadherin, and has been identified as an independent prognostic marker (139). Functional studies revealed that TMEM16A inhibition induced G0/G1 cell cycle arrest and reduced the invasiveness of breast cancer cells, underscoring its potential as a therapeutic target (139,140). However, conflicting data regarding its association with hormone receptors (HER2, ER and PR) suggested that the role of TMEM16A may vary across molecular subtypes and cellular contexts (140). Upregulation of TMEM16A is notably associated with CRC progression and higher TNM staging (141). Sui et al (142) confirmed that TMEM16A knockout inhibited the growth, migration and invasion of CRC cells by suppressing the MAPK/ERK signaling pathway. It has been reported that endogenous TMEM16A knockout suppressed CRC cell proliferation and induced apoptosis by inactivating the downstream Wnt/β-catenin signaling pathway (142). Li et al (143) proposed that TMEM16A mRNA expression was associated with lymph node metastasis in CRC and may thus serve as an independent predictive marker for lymphatic involvement. These findings establish the key role of TMEM16A in CRC progression and metastasis. Overexpression of TMEM16A is associated with TNM staging and Gleason scoring in prostate cancer, carrying out a key role in tumor growth, progression and metastasis (144).

Silencing or inhibiting endogenous TMEM16A has been revealed to suppress prostate cancer growth, induce apoptosis and increase TNF-α expression levels (145). Notably, natural products such as resveratrol and luteolin serve both as TMEM16A inhibitors and antitumor agents, effectively reducing prostate cancer cell proliferation and migration (146,147). These findings highlight the therapeutic potential of targeting TMEM16A in prostate cancer treatment. TMEM16A is considerably upregulated in epithelial OC cells, and its upregulation is associated with increased staging and lower differentiation (148). Downregulation of TMEM16A suppressed OC cell growth by reducing PI3K/Akt phosphorylation and disrupting the PI3K/Akt signaling pathway (148). These findings highlight the key role of TMEM16A in OC progression, with its expression associated with advanced disease and a poor prognosis. Given its role in cancer progression, TMEM16A has emerged as a promising therapeutic target. Inhibition of TMEM16A suppresses tumor growth, induces apoptosis and enhances the efficacy of existing therapies (144-149). For example, allicin, a natural compound from garlic, was shown to inhibit the chloride ion current through the TMEM16A channel and exhibit synergistic anticancer effects with cisplatin in lung cancer (149). Similarly, idebenone, a novel TMEM16A inhibitor, can block chloride channel activity and was shown to induce apoptosis in normal prostate and PC cells (150). Furthermore, TMEM16A inhibitors such as T16Ainh-A01 and CaCCinh-A01 have demonstrated efficacy in reducing tumor growth and invasiveness in prostate cancer (144).

TMEM16A, as a CaCC, carries out a key role in calcium-activated chloride secretion, which is involved in the pathogenesis of secretory diarrhea. Yin et al (151) revealed in rotavirus-induced diarrhea that viral enterotoxin NSP4 enhanced intracellular calcium levels, activating TMEM16A-mediated chloride secretion, which, in-turn, contributed to fluid loss and diarrhea. Given that chemotherapy-induced diarrhea in patients with cancer often involves dysregulated epithelial ion transport, TMEM16A may similarly contribute to this adverse effect. Thus, TMEM16A may serve as a mechanistic link between increased chloride secretion and diarrhea in both infectious and iatrogenic contexts.

TMEM16A, through its dual function as both a chloride channel and a signaling modulator, carries out a pivotal role in tumor biology. The regulation of cell proliferation, migration and metastasis by TMEM16A, in addition to its association with EGFR signaling, highlights its potential as both a prognostic biomarker and a therapeutic target for various malignancies.

TMEM45A

TMEM45A consists of 275 amino acids, composed of 5-7 transmembrane domains and is primarily localized in the Golgi apparatus (4). It is highly expressed in the skin, where it carries out a key role in epidermal differentiation and keratinization, processes essential for preserving the structural integrity of the skin and barrier function (152). Beyond its role in normal cellular functions, TMEM45A is involved in the pathogenesis of various types of cancer, including gliomas, breast cancer, HCC, RCC and OC, where it is commonly upregulated (153).

TGF-β is a key factor regulating various cellular processes such as proliferation, migration and invasion. In OC, the expression of TMEM45A is closely associated with the activation of the TGF-β signaling pathway. Silencing TMEM45A not only reduced cell proliferation, adhesion and invasion, but also downregulated cancer-promoting factors such as TGF-β1, TGF-β2, RhoA and ROCK2 (154). This association has been revealed to contribute to tumor progression and metastasis and a similar mechanism has been identified in glioma cell lines, providing further evidence for the oncogenic potential of TMEM45A in these cancer types (152,153). Notably, increased TMEM45A expression in breast cancer and cervical cancer (CC) has been associated with poor OS, indicating that TMEM45A could function as an important prognostic biomarker for these malignancies (155).

TMEM45A short hairpin RNA inhibited the proliferation of CC cells, arrested the cell cycle at the S phase and promoted apoptosis. It also suppressed EMT by downregulating Vimentin and N-cadherin, and upregulating E-cadherin, thereby reducing the invasive and migratory ability of the cells. The expression of TMEM45A in cisplatin-resistant cell lines SiHa/DDP and HeLa/DDP was revealed to be increased compared with their parental cells, and inhibiting TMEM45A markedly reduced the IC50 of these resistant cells to cisplatin. Silencing TMEM45A inhibited the proliferation, invasion, migration and EMT of HPV-positive CC cells, regulated cell cycle distribution and promoted apoptosis (156). Sun et al (153) revealed that TMEM45A was overexpressed in glioma tissues compared with non-tumorous brain tissues. TMEM45A mRNA levels progressively increased with higher histological grades of glioma and were associated with shorter survival times of patients. Knockout of TMEM45A in glioma cell lines markedly inhibited cell proliferation and induced G1 phase arrest, and reduced the migratory and invasive abilities of glioma cells. Notably, treatment with TMEM45A small interfering (si)RNA downregulated key proteins involved in cell cycle progression (Cyclin D1, CDK4 and proliferating cell nuclear antigen) as well as invasion-related proteins (MMP-2 and MMP-9), providing a potential mechanistic explanation for its role in glioma.

TMEM45A is closely associated with intratumoral heterogeneity, a major challenge in cancer treatment. In lung adenocarcinoma, TMEM45A was revealed to be overexpressed in tumor tissues compared with healthy tissues, and its expression was associated with markers such as GLUT1, MCT4 and CA9, which are associated with a hypoxic microenvironment and altered metabolism (157). Similarly, in ccRCC, TMEM45A upregulation associated with poor OS, disease-free survival and advanced clinicopathological features such as higher histological grade and TNM stage (158).

TMEM45A is a multifunctional protein associated with tumorigenesis, chemotherapy resistance and fibrosis. Its ability to modulate key signaling pathways, including TGF-β and Notch, positions it as a promising candidate for therapeutic targeting. In GC, increased TMEM45A expression was associated with poor prognosis and immune cell infiltration, making it a potential independent prognostic marker (159). In breast cancer, TMEM45A expression is associated with poor survival and resistance to CDK4/6 inhibitors such as palbociclib, further underscoring its prognostic value (160). Furthermore, the association of TMEM45A with a poor prognosis in various types of cancer and its involvement in MDR highlights its potential as a biomarker for cancer progression and therapeutic resistance (155,161,162). In breast cancer, engineered exosomes loaded with siRNA targeting TMEM45A restored sensitivity to palbociclib and suppressed tumor growth (160). Similarly, in CC, TMEM45A knockdown reversed cisplatin resistance and inhibited proliferation, migration and EMT in HPV-positive cell lines (156). Future research focusing on the molecular mechanisms underlying the functions of TMEM45A in these contexts are essential for the development of targeted therapies aimed at overcoming resistance and improving patient outcomes.

TMEM45B

TMEM45B has attracted considerable attention due to its involvement in various types of cancer, including lung cancer, PC and GC (163). Okada et al (163) have highlighted that TMEM45B as an oncogene that plays a pivotal role in cancer pathogenesis by regulating essential cellular processes, including proliferation, invasion, migration, and apoptosis.

In lung cancer, TMEM45B overexpression is associated with poor patient survival. Knockdown of TMEM45B reduced cell proliferation, induced cell cycle arrest and apoptosis, and inhibited cell migration and invasion. These effects were mediated through the regulation of cell cycle-related proteins (CDK2 and CDC25A), apoptosis-related proteins (Bcl2 and Bax) and metastasis-related proteins (MMP-9, Twist and Snail). TMEM45B may thus serve as a potential prognostic marker and therapeutic target in lung cancer (164). TMEM45B is highly expressed in PC tissues and cell lines. Its knockdown inhibited cell proliferation, invasion and migration while inducing cell cycle arrest and apoptosis. Conversely, TMEM45B overexpression promoted these processes and inhibited apoptosis. Gene set enrichment analysis revealed that TMEM45B regulated genes that are involved in the cell cycle and metastasis pathways, further supporting its oncogenic role in PC (165). Shen et al (166) also demonstrated that in GC, TMEM45B was overexpressed in both tissues and cell lines. Silencing TMEM45B inhibited cell proliferation, migration, invasion and EMT. This suppression was mediated through the inhibition of the JAK2/STAT3 signaling pathway, as evidenced by reduced levels of phosphorylated JAK2 and STAT3 upon TMEM45B knockdown (166).

Moreover, in osteosarcoma, Li et al (167) verified the upregulation of TMEM45B expression and revealed that its knockdown in U2OS cells effectively suppressed proliferation, migration and invasion in vitro, and inhibited tumor growth in xenograft models. This effect was associated with the downregulation of key proteins such as β-catenin, cyclin D1 and c-Myc, which are important for cancer cell proliferation and metastasis (167). These findings collectively indicate that TMEM45B is a potent regulator of tumorigenesis across various cancer types. Its dysregulated expression may act as a potential prognostic biomarker for cancer progression and it holds promise as a therapeutic target. Furthermore, previous studies have demonstrated that the regulatory effects of TMEM45B extend to modulating the TME, influencing immune responses and potentially modulating the response to chemotherapeutic agents (167,168). Similarly, TMEM45B has been identified as a novel predictive biomarker for prostate cancer progression and metastasis. Its expression is notably increased in tumor cell lines with high metastatic potential and is associated with biochemical recurrence, distant metastasis and a poor OS. Multivariate analysis confirmed that TMEM45B was an independent risk factor for metastasis, making it a promising prognostic marker for patients with prostate cancer (169).

TMEM45B exerts its oncogenic effects through multiple signaling pathways, including Wnt/β-catenin, JAK2/STAT3 and TGF-β. It regulates key cellular processes such as proliferation, apoptosis, migration and invasion by modulating the expression of downstream target genes and proteins. Additionally, the involvement of TMEM45B in EMT and chemotherapy resistance further underscores its role in cancer progression. Given its central role in regulating key pathways involved in cancer progression, TMEM45B may be developed as a targeted therapy, thereby improving patient outcomes.

TMEM48

TMEM48 is a member of the nuclear pore complex (NPC) family of proteins and is essential for maintaining nuclear membrane integrity and enabling nuclear transport (170). The protein consists of six transmembrane domains and is vital for NPC assembly, which is important for regulating several cellular processes, including chromosome segregation, gene transcription, DNA replication, DNA damage repair and apoptosis (170,171). Alterations in the expression of NPC proteins, including TMEM48, have been associated with the onset and progression of various malignancies, such as breast, melanoma, pancreatic, colon, gastric, prostate, esophageal, lung cancer and lymphoma (170-172).

Previous studies have highlighted the involvement of TMEM48 in NSCLC (170-173). Qiao et al (173) demonstrated that TMEM48 expression was markedly increased in NSCLC tissues compared with adjacent normal tissues. This overexpression was associated with a poor prognosis, including reduced survival times, increased tumor size and lymph node metastasis. TMEM48 regulates essential cellular processes, such as proliferation, migration and invasion, by regulating genes involved in the cell cycle and DNA replication. Knockdown of TMEM48 in NSCLC cell lines led to reduced cell proliferation, induction of G1 phase cell cycle arrest and inhibition of migration and invasion. Furthermore, silencing of TMEM48 triggered apoptosis in NSCLC cells. In an in vivo study, silencing TMEM48 led to a notable reduction in tumor weight in xenograft models (173).

Moreover, the study by Akkafa et al (6) suggests that microRNA (miR)-421, an miR targeting TMEM48, potentiates apoptotic signaling pathways in NSCLC. Suppression of TMEM48 by miR-421 upregulated the expression of apoptosis-related proteins, such as caspase-3, PTEN and p53, which collectively promote cell death. This indicated that targeting TMEM48 may serve as a therapeutic strategy for improving the efficacy of treatments in NSCLC. TMEM48 has been identified in other types of cancer, such as CC (174). In CC, TMEM48 is overexpressed and its knockdown inhibited cell proliferation, migration and invasion both in vitro and in vivo (174). Jiang et al (174) indicated that TMEM48 promotes CC progression by activating the Wnt/β-catenin signaling pathway, as evidenced by the reduced expression of β-catenin, TCF1 and AXIN2 following TMEM48 knockdown. This suggested that TMEM48 is important in tumorigenesis by modulating key signaling pathways involved in cancer progression.

TMEM48 is a promising prognostic marker and therapeutic target. Given its central role in regulating cellular functions key for tumor progression, TMEM48 represents a valuable target for future therapeutic strategies focused on inhibiting cancer growth and metastasis. Further research is needed to explore the therapeutic potential of TMEM48 in various malignancies, including its modulation by miRNAs.

TMEM65

TMEM65 is a mitochondrial inner membrane protein whose deficiency has been implicated in a mitochondrial disorder characterized by a complex encephalomyopathic phenotype (175,176). Clinical manifestations include microcephaly, craniofacial dysmorphism, psychomotor regression, generalized hypotonia, growth retardation, lactic acidosis, intractable seizures and dyskinetic movements, notably in the absence of cardiomyopathy (175). Previously, TMEM65 has emerged as a pivotal factor in cancer biology, with its involvement reported across various malignancies, including CRC, TNBC, GC and HCC (175-180). TMEM65 carries out a multifaceted role in tumorigenesis by modulating mitochondrial dynamics, driving metabolic reprogramming and regulating signaling pathways associated with cancer progression and therapeutic resistance. This contrast between its role in congenital mitochondrial disease and acquired malignancies underscores its biological value and potential as a therapeutic target.

In CRC, TMEM65 is transcriptionally regulated by the CHD6-TCF4 axis, which is activated by both EGF and Wnt signaling pathways. CHD6, a chromatin remodeler, binds to Wnt signaling transcription factor TCF4 to facilitate TMEM65 expression. This axis promotes mitochondrial dynamics and ATP production, essential for cancer cell proliferation, migration and invasion. Targeting the CHD6-TMEM65 axis with Wnt inhibitors and anti-EGFR therapies showed promise in restricting CRC growth (176). TMEM65 acts as an oncogene in TNBC, where it was revealed to be upregulated by the transcription factor MYC and DNA demethylase TET3. TMEM65 enhanced mitochondrial oxidative phosphorylation and reactive oxygen species production, which, in turn, activates hypoxia inducible factor (HIF)-1α and SERPINB3 and promotes cancer stemness, progression, and cisplatin resistance. Inhibition of MYC and TET3 attenuated TMEM65-driven TNBC progression, highlighting its potential as a therapeutic target (177). TMEM65 amplification and overexpression are common in GC and are associated with a poor prognosis. TMEM65 promoted tumorigenesis by activating the PI3K-Akt-mTOR signaling pathway through its interaction with YWHAZ, a protein that inhibits ubiquitin-mediated degradation. Silencing TMEM65 suppresses tumor growth and metastasis, making it a promising therapeutic target in GC (163). In HBV-related HCC, TMEM65 amplification was driven by HBV integration-induced genomic rearrangements. TMEM65 contributes to tumorigenesis by promoting mitochondrial function and metabolic reprogramming, which are key for cancer cell survival and proliferation (178).

TMEM65 expression is associated with tumor immune infiltration, particularly CD8+ T effector cells and immune checkpoint markers. Its role in modulating the TME and immune response makes it a potential biomarker for predicting prognosis and the efficacy of immunotherapy (179). TMEM65 may serve as a target for chimeric antigen receptor T-cell therapies and antibody-drug conjugates. Its overexpression in cancer tissues and low expression in normal tissues make it an attractive candidate for immune-based interventions (180). TMEM65 is a pivotal oncogene that drives cancer progression through its roles in mitochondrial dynamics, metabolic reprogramming and signaling pathway activation. Its overexpression is associated with poor prognosis across multiple cancer types, making it a promising biomarker and therapeutic target. Future research should focus on developing targeted therapies that exploit the oncogenic function of TMEM65 while minimizing off-target effects.

TMEM140

TMEM140, also known as FLJ11000, is a TP encoded by a gene located on chromosome 7q33, consisting of 185 amino acids (181). Although initially identified for its role in supporting hematopoietic stem cells in vitro (181-183), TMEM140 has since been associated with various biological processes, including tumorigenesis and viral infection (182). Guan et al (183) have highlighted the dual functions of TMEM140 as both a prognostic biomarker and a potential therapeutic target in cancer and its role in inhibiting herpes simplex virus 1 (HSV-1) replication. TMEM140 has emerged as a key player in various types of cancer, including gliomas and GC. Gliomas are among the most common and aggressive primary brain tumors, characterized by a poor prognosis and limited treatment options. TMEM140 is overexpressed in gliomas compared with normal brain tissues and its high expression is associated with a larger tumor size, higher histologic grade and shorter OS (181). An In vitro study using glioma cell lines have revealed that silencing TMEM140 inhibited cell proliferation, migration and invasion, with concurrent arrest in the G1 phase of the cell cycle and activation of apoptotic pathways (181). It regulates cell cycle progression by shortening the cell cycle and reducing apoptosis, thereby promoting tumor cell proliferation. Additionally, TMEM140 enhanced tumor cell adhesion and survival by modulating the expression of adhesion and anti-apoptotic proteins (181-183). In vivo silencing of TMEM140 led to a marked reduction in tumor volume and weight (183), highlighting its pivotal role in tumor growth. These findings identify TMEM140 as a novel prognostic biomarker and therapeutic target for gliomas, underscoring its potential for clinical application in glioma treatment. Beyond gliomas, TMEM140 has been implicated in other cancer types through its co-expression with long-chain acyl-coenzyme A synthetase 5 and this was predictive of improved prognosis in breast, ovarian and lung cancer (184). This suggests that TMEM140 may carry out a context-dependent role in cancer progression, potentially acting as a tumor suppressor in certain malignancies (184). Furthermore, TMEM140 has been identified as a candidate gene in HTLV-1-infected cancer cells, indicating its broader relevance in cancer biology (185).

TMEM140 is a multifunctional protein that is involved in several processes in cancer biology, autoimmune diseases and viral infections. In systemic sclerosis, the involvement of TMEM140 in DNA methylation changes indicates a potential key role in disease pathogenesis, particularly in fibroblast function (186). Additionally, its role in inhibiting HSV-1 replication highlights its promise as an antiviral target. Given its diverse functions, TMEM140 may serve as a compelling candidate for further investigation, with implications for therapeutic strategies across a variety of diseases.

TMEM158

TMEM158, located on chromosome 3p21.3, also known as BBP, Ras-induced senescence 1 protein, p40BBP and HBB, was initially recognized as an upregulated potential tumor suppressor gene in the context of Ras V12 lentiviral infection-induced senescence in fibroblasts (187). Li et al (188) have highlighted its key role in the development of various types of cancer. Importantly, TMEM158 is overexpressed in Wilms tumor (nephroblastoma), with somatic mutations in the β-catenin gene, indicating a connection between Ras and Wnt signaling pathways (189). In lung cancer, TMEM158 is highly expressed in advanced-stage tumors and is associated with poor OS (61). It promotes EMT and enhanced cell migration, contributing to aggressive tumor behavior (61). Additionally, TMEM158 expression is induced under hypoxic conditions in a HIF-1α-dependent manner, associating it with tumor hypoxia, a hallmark of cancer progression (61). Moreover, TMEM158 has emerged as a potential biomarker for predicting the efficacy of cisplatin-based therapy in NSCLC, with its expression correlating with improved therapeutic outcomes (61,190).

In PC, TMEM158 overexpression promotes cell proliferation, migration and invasion through the activation of the TGFβ1 and PI3K/AKT signaling pathways, highlighting its oncogenic potential (191). TMEM158 is amplified in GC tissues and is associated with poor prognosis (192). In GC, it was revealed to accelerate GC cell proliferation by activating the PI3K/AKT signaling pathway and its knockdown inhibited tumor growth, suggesting its potential as a biomarker and therapeutic target (192). In CRC, downregulation of TMEM158 expression notably impairs tumor cell proliferation, migration and MDR, highlighting its value in both tumor progression and chemoresistance (193). Furthermore, TMEM158 is involved in regulating EMT in various types of cancer, enhancing metastasis and aggressiveness. In OC, TMEM158 expression was shown to be considerably upregulated, thereby promoting tumorigenic properties, including cell proliferation, invasion, adhesion and migration (194). Silencing TMEM158 resulted in downregulation of intercellular adhesion molecules (ICAM1 and VCAM1) and disruption of the TGF-β signaling pathway, further emphasizing its role as an oncogene (194). It also contributed to doxorubicin resistance by upregulating ABCG2, a drug efflux transporter (195). TMEM158 is highly expressed in TNBC and was shown to be associated with poor clinical outcomes and to drive EMT and tumor metastasis via activation of the TGF-β signaling pathway, highlighting it as a potential therapeutic target for this aggressive breast cancer subtype (196).

In glioblastoma (GBM), overexpression of TMEM158 was associated with a poor prognosis, whereas silencing it suppressed cell migration and invasion by interfering with STAT3 signaling pathways (188). Unlike its oncogenic role in other types of cancer, TMEM158 is downregulated in prostate cancer and is associated with advanced disease features and poor survival. Its expression is negatively regulated by androgen receptor signaling, and it may carry out a tumor-suppressive role in this context (197,198). Additionally, TMEM158 was shown to be involved in immune modulation, correlating with tumor mutational burden, microsatellite instability and immune cell infiltration in the TME (190,197,198).

In NSCLC, TMEM158 is considered to modulate immune checkpoints, thereby enhancing the therapeutic efficacy of immune checkpoint inhibitors such as anti-PD-1 and anti-CTLA-4 (190,197,198). This emerging role in immuno-oncology highlights a promising avenue for therapeutic interventions targeting TMEM158. Additionally, pan-cancer analyses demonstrated that TMEM158 is upregulated in a range of cancer types, including melanoma and HCC, where its elevated expression is associated with tumor aggressiveness and poor prognosis (197). In these contexts, TMEM158 has been implicated in hypoxia-induced pathways, thus facilitating tumor growth in hypoxic conditions. These findings collectively highlight the key role of TMEM158 in regulating tumor progression, metastasis, immune evasion and chemoresistance across various cancer types. Its involvement in signaling pathways, including TGFβ, PI3K/AKT, STAT3 and Wnt, as well as its potential in immune modulation, establishes TMEM158 as both a prognostic biomarker and a promising therapeutic target in oncology.

TMEM159

TMEM159, also known as promethin, is a small TP consisting of 161 amino acids with four adjacent transmembrane helices forming two helical hairpins (199). It carries out a key role in lipid droplet (LD) biogenesis through its interaction with Seipin, a key regulator of lipid storage in the ER (200). The TMEM159-Seipin complex co-purified with triacylglycerol and facilitated the initiation of LD formation. Upon maturation of LD, TMEM159 dissociated and relocated to the LD surface, thereby participating in the maintenance of lipid homeostasis (200).

Although extensively studied in the context of lipid metabolism, TMEM159 has recently gained attention for its potential involvement in cancer progression (199-202). Notably, in GBM, enhanced LD formation, which is associated with a poor prognosis under metabolic stress, has been associated with altered TMEM159 expression (201). TMEM159 may contribute to glioma malignancy by modulating EGFR signaling pathways, which are important for cancer cell proliferation, migration and survival (201). These findings suggest that TMEM159 could serve as both a molecular biomarker and therapeutic target in lipid-associated malignancies such as GBM (201,202).

Beyond cancer, TMEM159 has also been associated with neuropsychiatric disorders through genome-wide association studies (199-205), but its mechanistic involvement in tumorigenesis, particularly through pathways associated with lipid regulation and oncogenic signaling, may represent promising directions for future cancer-focused research. Continued investigation is warranted to fully elucidate the multifaceted role of TMEM159 in lipid metabolism and cancer biology (203-205).

TMEM205

TMEM205 is a TP encoded by a gene located on chromosome 19p13.2, and is formed of 189 amino acids. TMEM205 has emerged as a notable player in cancer biology, particularly in the context of drug resistance, tumor progression and immune modulation. TMEM205 has shown promise as a biomarker for chemoresistance and prognosis. In high-grade serous OC, TMEM205 was found to be differentially expressed in extracellular vesicles derived from platinum-resistant patients with OC, demonstrating high diagnostic accuracy for platinum resistance (206). In HCC, low TMEM205 expression was independently associated with poor OS and disease-specific survival, highlighting it as a potential prognostic marker (207). TMEM205 is associated with chemoresistance, particularly to platinum-based therapies such as cisplatin (208). Moreover, TMEM205 was also shown to carry out a role in tumor progression and immune evasion (208). Targeting TMEM205 may offer a novel strategy to overcome chemoresistance and improve treatment outcomes. In ovarian clear cell carcinoma, the combination of oncolytic viruses and cisplatin reduced TMEM205 expression and sensitized cells to chemotherapy (208). Similarly, small molecule inhibitors such as L-2663 have shown efficacy in enhancing platinum sensitivity in OC (209). These findings underscore the potential of TMEM205 as a therapeutic target in combination therapies.

TMEM205 is a multifaceted protein involved in chemoresistance, tumor progression and immune modulation across various types of cancer. Its upregulation is associated with poor treatment outcomes, while its inhibition enhances chemotherapy efficacy. As a biomarker, TMEM205 may serve as a diagnostic and prognostic value, particularly in various types of platinum-resistant cancer (206-209). Future research should focus on developing targeted therapies to modulate TMEM205 activity, potentially improving patient outcomes in various types of resistant cancer.

TMEM proteins involved in TME

The TME, composed of immune cells, stromal components, extracellular matrix and soluble factors, carries out a pivotal role in cancer progression, immune evasion and therapy response. Growing evidence shows that TPs not only regulate tumor cell behavior but also engage in bidirectional crosstalk with the TME, contributing to its remodeling and tumor development (4,5).

Several TMEM proteins have been shown to regulate tumor-immune interactions. For example, TMEM173 carries out a key role in the activation of innate immune responses and can modulate T-cell infiltration and antitumor immunity within the TME (210-212). Evidence from several types of cancer, including SCLC, TNBC, CC and GBM, demonstrates that activation of the cGAS-STING pathway can reshape the TME by promoting chemokine production (such as CXCL10 and CCL5), increasing CD8+ T cell infiltration and potentiating antitumor immunity (210,212-214). Microbiota-derived STING agonists can reprogram monocytes and DC in the TME, enhancing IFN-I production and improving the efficacy of immune checkpoint blockade (215). Moreover, STING signaling can be triggered not only by endogenous cytosolic DNA but also by tumor-derived exosomes, indicating a dynamic extracellular-intracellular feedback loop that facilitates immune regulation (213). STING was also shown to contribute to immunosuppression in certain contexts, such as by promoting regulatory T cell (Treg) differentiation or inducing ER stress and T cell apoptosis through non-canonical pathways (213,216). Senescent tumor cells release mitochondrial DNA that activates the cGAS-STING pathway in myeloid-derived suppressor cells, promoting immunosuppression and tumor progression (107). These multifaceted roles highlight TMEM173 as a key modulator of immune surveillance and tumor progression, reflecting the complex bidirectional communication between TMEM proteins and the TME.

TMEM205 carries out a key role in modulating TME and influencing cancer progression. In HCC, TMEM205 expression was shown to be associated with the immune landscape of the tumor, where low TMEM205 expression was associated with a poor prognosis and survival outcomes (207,209). TMEM205 interacted with various immune cell populations in the TME, particularly macrophages and Tregs (207,209). The expression of TMEM205 was positively associated with the proportion of macrophages, including M1 macrophages, while it was negatively associated with immunosuppressive M2 macrophage markers and Treg markers. This suggested that TMEM205 may exert its antitumor effects by reducing the levels of immunosuppressive cells such as M2 macrophages and Tregs, while promoting the infiltration of cytotoxic CD8+ T cells into the TME. Therefore, extracellular feedback through TMEM205 may regulate immune cell infiltration, enhancing antitumor immunity and potentially improving therapeutic responses, especially in combination therapies targeting immune checkpoints (207,209).

TMEM16A also carries out a key role in shaping the TME. In PC, high TMEM16A expression is associated with an immunosuppressive TME, characterized by increased cancer-associated fibroblasts (CAFs) and reduced CD8+ T cell infiltration (217). Similarly, in gastrointestinal cancer, TMEM16A contributed to immunotherapy resistance by inhibiting cancer ferroptosis and recruiting CAFs, thereby impairing CD8+ T cell-mediated anti-tumor immunity (218). TMEM101 also carried out a key role in modulating the HCC TME and influencing HCC progression. In HCC, TMEM101 expression was markedly upregulated, and was associated with a poor prognosis and advanced tumor grade and stage (219). TMEM101 was shown to interact with various immune cell populations within the TME, particularly influencing macrophage infiltration and T cell activity. High TMEM101 expression is associated with increased M0 macrophage infiltration, which are pro-tumorigenic and decreased infiltration of anti-tumor immune cells such as M1 macrophages and resting memory CD4+ T cells. This shift in immune cell populations suggests that TMEM101 promotes tumor progression by modifying the immune landscape in favor of tumor growth. Furthermore, intracellular feedback regulation by TMEM101 was shown to impact pathways related to cell cycle, DNA replication and repair, facilitating tumor proliferation and survival (219). Thus, TMEM101 not only serves as a diagnostic and prognostic biomarker for HCC but also acts as an oncogene that influences the TME, contributing to HCC initiation and progression (219).

TMEM proteins also mediate extracellular and intracellular feedback regulation. Extracellularly, they can interact with cytokines, growth factors and adhesion molecules, modulating upstream signaling events. TMEM52B is a novel regulator that modulates the activity of both E-cadherin and EGFR, exhibiting tumor suppressor-like functions. Lee et al (220) revealed that peptides derived from the extracellular domain (ECD) of TMEM52B considerably inhibit cancer cell survival, invasion and anchorage-independent growth, while reducing the phosphorylation of ERK1/2 and AKT. These peptides stabilize E-cadherin at cell-cell junctions, leading to decreased β-catenin activity, and directly bind to the E-cadherin ECD to block the interaction between soluble E-cadherin and EGFR, thereby suppressing EGFR activation. These findings suggest that TMEM52B ECD-derived peptides not only possess strong anti-cancer potential but also serve as valuable tools for investigating the regulatory role of TMEM52B in the crosstalk between E-cadherin and EGFR (220). Intracellularly, TMEM proteins often participate in key oncogenic pathways such as MAPK, PI3K/AKT and Wnt, forming regulatory loops that either amplify or attenuate cellular responses to environmental cues. These feedback mechanisms may affect cell proliferation, apoptosis, migration and ultimately, cancer progression (4,5,220).

Collectively, the bidirectional interaction between TMEM proteins and the TME contributes to tumor plasticity, immune evasion and chemoresistance. Although this area remains underexplored, it represents a promising direction for future research. An improved understanding of how TMEM proteins influence the TME may lead to the identification of novel therapeutic targets and strategies to modulate tumor-host interactions.

TMEM proteins in tumor chemoresistance

TMEM45A

Hypoxia is a defining characteristic of solid tumors, markedly contributing to resistance to both radiotherapy and chemotherapy, which increases the likelihood of tumor recurrence (221,222). This phenomenon is particularly prominent in liver and breast cancer, where the hypoxic TME provides a protective effect against chemotherapy-induced cell death. In these conditions, TMEM45A has been shown to serve as a key mediator of tumor progression and resistance to chemotherapeutic agents under hypoxic conditions (156-161,221,222). Previous studies have demonstrated an upregulation of TMEM45A in hypoxic tumor cells, such as HepG2 and MDA-MB-231 cells, exposed to chemotherapeutic agents such as Taxol or etoposide. Under these conditions, TMEM45A protects against apoptosis, a key mechanism that drives chemotherapy resistance (156-161,221,222). Specifically, silencing TMEM45A using siRNA alleviated the protective effect of hypoxia and sensitized the cells to chemotherapy-induced cell death. This indicates that TMEM45A carries out a key role in sustaining chemoresistance under hypoxic stress (156-161,221,222). The role of TMEM45A in modulating chemosensitivity depends on the cancer type, with its activation potentially acting as either a pro-apoptotic or anti-apoptotic factor.

In HNSCC and RCC, TMEM45A is upregulated, and its upregulation is associated with a poorer prognosis. In these types of cancer, silencing TMEM45A enhanced sensitivity to cisplatin, a commonly used chemotherapeutic agent, by modulating DNA damage repair mechanisms and the unfolded protein response pathway (156-161,221,222). Conversely, in breast and liver cancer cells, TMEM45A contributed to chemoresistance by inhibiting apoptotic pathways, thus preventing cell death induced by Taxol and etoposide (156-161,221,222). Further investigation into the mechanisms by which TMEM45A regulates chemoresistance uncovered its involvement in regulating the EMT, a process associated with enhanced tumor invasion and metastasis (156-161,221,222). In CC, silencing TMEM45A downregulated EMT markers such as Vimentin and N-cadherin and increased the expression of the epithelial marker E-cadherin. This indicated that TMEM45A not only influenced chemoresistance but also promoted tumor progression through EMT, thereby enhancing the invasive potential of cancer cells (156-161,221,222). The association between TMEM45A expression and chemotherapy resistance has been demonstrated in HPV-positive CC cells. In this context, TMEM45A was highly expressed in SiHa and HeLa cells and silencing TMEM45A notably reduced cisplatin resistance, as evidenced by a decrease in the IC50 values of cisplatin-treated SiHa/DDP and HeLa/DDP cells. This finding highlights the potential of TMEM45A as a therapeutic target for overcoming chemoresistance in HPV-positive CC (156-161,221,222).

TMEM45A expression is upregulated in Palbociclib-resistant HR+ breast cancer cells and is associated with accelerated tumor progression and increased glycolysis (156-161,221,222). Its overexpression drove Palbociclib resistance, while silencing TMEM45A enhanced sensitivity to the drug, induced cell cycle arrest and apoptosis, and inhibited cancer cell proliferation. Mechanistic analysis revealed that TMEM45A downregulated EMT markers and glycolysis-related proteins, and this reduced tumor invasiveness. By activating the AKT/mTOR pathway, TMEM45A regulated the cell cycle and glycolysis (4,5,156-161,221,222). In mouse models, TMEM45A knockdown restored Palbociclib sensitivity and inhibited tumor growth (156-161,221,222). Exosome-mediated delivery of TMEM45A-targeting siRNA in patient-derived xenografts enhanced sensitivity to CDK4/6 inhibitors without toxicity (160).

Furthermore, TMEM45A carries out a key role in MDR, especially in CRC. Studies have shown that TMEM45A expression is upregulated in 5-fluorouracil (5-FU)-resistant CRC cells (155-162). Knockdown of TMEM45A increased the sensitivity of these cells to 5-FU, induced apoptosis and increased the intracellular accumulation of drugs. Additionally, inhibition of TMEM45A suppressed cell migration and invasion and reduced the activation of the TGF-β signaling pathway, which is known to promote EMT and contribute to drug resistance in cancer (155-162). These findings underscore the potential of TMEM45A as a target for overcoming MDR in CRC.

In conclusion, TMEM45A is a key regulator of chemoresistance in various types of cancer, particularly when a hypoxic environment is present. Its dual role in promoting either pro-apoptotic or anti-apoptotic pathways, which varies depending on the cancer type, highlights its complexity as a therapeutic target. Targeting TMEM45A may, therefore, serve as a promising strategy to enhance the efficacy of chemotherapy, especially in hypoxic tumors prone to resistance. Further studies are needed to fully elucidate the molecular pathways through which TMEM45A exerts its effects on tumor progression and drug resistance, offering key insights into the development of novel therapeutic strategies (156-162,221,222).

TMEM158

TMEM158 carries out a key role in mediating chemoresistance in several types of cancer, including CRC. Overexpression of TMEM158 increases resistance to cisplatin and 5-FU by upregulating MDR-related proteins such as MDR1, MRP1 and Bcl-2, thereby enhancing drug efflux and inhibiting apoptosis (188,190-194). Silencing TMEM158 decreased the levels of these MDR proteins, restoring drug sensitivity. Conversely, its overexpression in drug-resistant CRC cells increased the IC50 values for cisplatin and 5-FU, highlighting its role in chemoresistance (188,190-194). Modulation of the intrinsic apoptotic pathway by TMEM158 further contributed to its impact on tumor survival and drug resistance (193). In cisplatin-resistant NSCLC PC-9/CDDP cell lines, knockdown of TMEM158 enhanced the sensitivity of the cells to cisplatin, demonstrating that TMEM158 carries out a key role in the development of chemoresistance (194). Moreover, TMEM158 expression is not only associated with resistance to cisplatin in NSCLC but also acts as a predictive biomarker for treatment response. Given its role in regulating tumor progression and drug resistance, TMEM158 may serve as a potential biomarker for predicting treatment response and as a potential therapeutic target for overcoming chemoresistance.

TMEM88

TMEM88 is a dual-transmembrane protein implicated in several types of cancer, including breast, ovarian, lung and testicular cancer. Research suggests that TMEM88 carries out a key role in mediating tumor progression and is notably associated with chemoresistance. TMEM88 modulates essential cellular processes, including cell proliferation, differentiation and apoptosis, by regulating the Wnt signaling pathway (74-77,223,224).

In OC, TMEM88 expression is associated with cisplatin resistance, a commonly used chemotherapeutic agent. Studies revealed that in cisplatin-resistant OC cells, the promoter methylation of TMEM88 was markedly reduced, leading to its overexpression (223). TMEM88 inhibited the canonical Wnt signaling pathway and its overexpression or knockdown affected tumor cell sensitivity to chemotherapy. Silencing TMEM88 in resistant cells enhanced sensitivity to cisplatin and increased the expression of Wnt target genes, such as cyclin D1 and c-Myc, further confirming its role in modulating Wnt signaling (75,223). Treatment with DNA methyltransferase inhibitors, such as SGI-110, restored TMEM88 expression, reversing chemoresistance and increasing cellular sensitivity to cisplatin (223,224). TMEM88 inhibited canonical Wnt signaling by interacting with DVL, a key mediator in the pathway. This inhibition carries out a dual role, suppressing tumor progression and influencing chemoresistance by modulating the TME.

TMEM88 has emerged as a promising therapeutic target for overcoming chemoresistance across several types of cancer. Targeting TMEM88 through epigenetic modulation or gene therapy may reverse chemoresistance and enhance chemotherapy efficacy. The role of TMEM88 in the Wnt signaling pathway offers a molecular basis for its involvement in tumor progression and drug resistance, highlighting novel avenues for therapeutic strategies in the management of cancer. Further studies are required to fully elucidate the mechanisms underlying the role of TMEM88 in tumor progression and chemoresistance, which may lead to the development of novel therapeutic interventions to restore chemotherapy sensitivity (74-82,223,224).

TMEM205

TMEM205 has been identified as a key player in cisplatin resistance across several types of cancer, including OC and GC (207-209). It mediates resistance through various mechanisms, particularly by enhancing drug efflux via the exosomal pathway. TMEM205 has been associated with increased exosome release, which can expel cisplatin from cancer cells, reducing its intracellular accumulation and thus contributing to resistance (209,225,226).

In OC, TMEM205 was revealed to be upregulated in cisplatin-resistant cell lines, and its inhibition using specific small molecules or genetic knockdown notably increased cisplatin accumulation in these cells, enhancing their sensitivity to the drug (208,227). Additionally, TMEM205 was shown to co-localize with RAB8, a protein involved in vesicular trafficking, further suggesting its role in the regulation of exosome-mediated drug efflux (208,227). Combining TMEM205 inhibition with cisplatin or oncolytic virus therapy was found to restore chemosensitivity and inhibit tumor growth (209,226).

In GC, TMEM205 also carries out a key role in promoting cisplatin resistance by inducing M2 polarization of TAMs, which enhanced cancer cell stemness, migration and EMT. These processes contribute to the malignant progression of the disease (208,225). The modulation of TMEM205 expression can alter the immune environment of the tumor, further influencing chemoresistance (227).

Overall, TMEM205 represents a potential therapeutic target to overcome cisplatin resistance, and strategies such as its inhibition, combined with other treatments, may improve the efficacy of chemotherapy in different types of resistant cancer.

TMEM16A

Recent research has highlighted the key role of TMEM16A in mediating chemoresistance, particularly to platinum-based therapies in HNSCC (228). TMEM16A overexpression has been revealed to increase oxidative stress, which upregulates ATP7B, a copper-transporting ATPase that facilitates lysosomal sequestration of platinum compounds, such as cisplatin, thereby diminishing their cytoplasmic efficacy. This mechanism underpins a novel resistance pathway, as demonstrated by a positive association between TMEM16A and ATP7B mRNA levels in squamous cell carcinoma of the head and neck tumor specimens (133). Notably, interventions with copper chelators (such as cuprizone and bathocuproine disulfonic acid) and antioxidants (for example, N-acetylcysteine) effectively reversed ATP7B induction and restored cisplatin sensitivity, especially in TMEM16A-overexpressing cells (229). These findings suggest that TMEM16A contributes to a copper-dependent resistance mechanism, offering potential therapeutic implications. Additionally, TMEM16A expression may serve as a biomarker to guide the selection of chemotherapy. For example, patients with cisplatin-resistant oral squamous cell carcinoma in high-risk groups showed enhanced sensitivity to docetaxel and shikonin, indicating the potential of TMEM16A in risk stratification and treatment optimization (229).

Discussion

The present review discusses the roles of various TMEM proteins in tumorigenesis and their potential contributions to chemoresistance, providing insights into their mechanisms and potential therapeutic implications (Table I). TMEM proteins have been implicated in multiple types of cancer, in which they regulate fundamental cellular processes such as proliferation, invasion, migration and apoptosis. The ability of TMEM proteins to influence tumor progression and resistance to therapy highlights their potential as targets for cancer treatment (Fig. 4). TMEM proteins carry out a key role in regulating cancer cell proliferation, migration and invasion through various molecular mechanisms and signaling pathways, which can be categorized into four key aspects (Table I and Fig. 4).

Roles and functions of TMEMs in
cancer development, immune response and cellular regulation,
providing a view of their functions across different pathways.
TMEM, transmembrane; EMT, epithelial-mesenchymal transition; DVL,
Dishevelled.

Figure 4

Roles and functions of TMEMs in cancer development, immune response and cellular regulation, providing a view of their functions across different pathways. TMEM, transmembrane; EMT, epithelial-mesenchymal transition; DVL, Dishevelled.

Table I

List of TMEM proteins involved in cancer.

Table I

List of TMEM proteins involved in cancer.

NameAmino acidsLocationInvolvement in cancerMain functionMechanism(Refs.)
TMEM7232Chromosome 3p21.3Tumor suppressor; involved in liver cancer.Inhibits the proliferation, migration, and carcinogenesis of HCC cells.Mediates epigenetic silencing through DNA methylation and histone deacetylation.(66-69)
TMEM16A986Chromosome 11q13.3Oncogene; involved in breast, esophageal, head and neck cancers.Chloride channel; regulates cell proliferation, migration and EMT.Interacts with EGFR signaling pathways; involved in the activation of EGFR and CAMK pathways, promoting metastasis through EMT.(126-151, 217,218, 228,229)
TMEM25366Chromosome 11q23.3Tumor suppressor, involved in colon, breast, and renal cell carcinoma.Modulates immune response; regulates growth factor signaling; affects cell adhesion; suppresses tumor progression.Mediates epigenetic silencing through DNA methylation and histone deacetylation; inhibits EGFR-mediated STAT3 activation; facilitates immune evasion.(12,70-73)
TMEM45A275Chromosome 3q12.1Oncogene; involved in gliomas, breast, liver, renal, ovarian cancer.Mediator of chemoresistance under hypoxic and drug-stressed conditions; involved in epidermal differentiation and keratinization.Activates the TGF-β pathway; enhances chemoresistance in hypoxic conditions and metabolic adaptation; promotes EMT; regulates the cell cycle and apoptosis.(4,152-162,221, 222)
TMEM45B275Chromosome 11q24.3Oncogene; involved in lung, pancreatic, gastric cancer.Promotes cell proliferation, migration, invasion; inhibits apoptosis.Regulates STAT3/JAK2 pathways; modulates tumor microenvironment by influencing immune responses and hypoxic adaptation.(163-169)
TMEM48674Chromosome 1p32.3Oncogene; involved in NSCLC, cervical cancer.Regulates nuclear transport and membrane integrity; promotes tumor cell proliferation, migration, and invasion.Activates the Wnt/β-catenin signaling pathway; regulates the cell cycle and apoptosis.(170-174)
TMEM65240Chromosome 8q24.1Oncogene; involved in TNBC, HCC, CRC and GCRegulates mitochondrial dynamics and OXPHOS; promotes cell proliferation and stemness; modulates immune response and tumor microenvironment.Transcriptionally activated by CHD6-TCF4 in CRC, by MYC and TET3 in TNBC; Activates PI3K-Akt-mTOR pathway via interaction with YWHAZ in GC; enhances ROS-HIF-1α-SERPINB3 axis to promote stemness and cisplatin resistance in TNBC.(175-180)
TMEM88159Chromosome 17p13.1Tumor suppressor; involved in breast, ovarian, lung, thyroid cancer.Inhibits Wnt signaling and regulates chemoresistance; involved in cardiomyocyte differentiation.Epigenetic modulation; inhibits the Wnt pathway by directly interacting with DVL; subcellular localization-dependent functional switch.(74-82, 223,224)
TMEM106A262Chromosome 17q21.3Tumor suppressor; involved in gastric and liver cancer.Immune modulation; induces apoptosis, delays tumor growth, activates the caspase pathway.Promoter methylation; caspase activation; inhibition of EMT; inactivation of the ERK1/2/Slug pathway; regulation of macrophage activation and inflammatory responses.(83-91)
TMEM140185Chromosome 7q33Oncogene; involved in gliomas, gastric cancer.Supports glioma cell proliferation, migration and invasion; enhances tumor cell adhesion and survival.Promotes tumor aggressiveness and metastasis by regulating the cell cycle (G1 arrest) and inhibiting apoptosis, enhances expression of adhesion and anti-apoptotic proteins.(181-186)
TMEM158300Chromosome 3p21.3Oncogene; involved in NSCLC, pancreatic, colorectal, ovarian cancer.Promotes cell proliferation and survival; regulates EMT.Activation of TGF-β, PI3K/AKT and STAT3 pathways; implicated in Ras signaling and involved in tumor progression; regulates immune responses.(61,187-198)
TMEM159161Chromosome 16p12.3Oncogene; involved in glioblastomaRegulates lipid droplet biogenesis, associated wth lipid metabolism; promotes cell proliferation and survival.Interacts with Seipin; modulates lipid metabolism; implicated in psychiatric disorders; affects EGFR pathways.(199-205)
TMEM173379Chromosome 5q31.2Tumor suppressor; involved in various types of cancer including gastric and liver cancer.Regulates type I interferon signaling, innate immunity, antitumor immunity.Modulates autophagy and oxidative stress response; involved in T cell activation, immune checkpoint modulation, and macrophage polarization.(92-108, 210-216)
TMEM176A235Chromosome 7q36.1Tumor suppressor; involved in different types of cancer such as ESCC and CRC, linked with promoter methylation.Suppresses tumor growth, regulates immune responses.Acts through promoter methylation; inhibits ERK1/2 pathways; modulates immune responses.(111-113)
TMEM176B270Chromosome 7q36.1Tumor suppressor; involved in lung adenocarcinoma and GC progression.Modulates immune tolerance and antigen cross-presentation; regulates immune checkpoint responses.Regulates phagosomal pH, induces CD8+ T cell activation, interacts with immune checkpoints, and may serve as a predictor of immune checkpoint inhibitor response; facilitates immune evasion; involves tumor-associated macrophages.(114-123)
TMEM205189Chromosome 19p13.2Oncogene; involved in GC, OC and HCC.Mediates chemoresistance, tumor growth, and immune modulation.Involves the exosomal pathway and drug efflux; modulates macrophage polarization and immune evasion; promotes immune cell infiltration and cytotoxic T cell activity.(206-209, 225-227)

[i] CAMK, Calcium/calmodulin-dependent protein kinase; CHD6, chromodomain helicase DNA binding protein 6; CRC, colorectal cancer; DVL, Dishevelled; EGFR, epidermal growth factor receptors; EMT, epithelial-mesenchymal transition; ERK, extracellular regulated protein kinases; ESCC, esophageal squamous cell carcinoma; GC, gastric cancer; HCC, hepatocellular carcinoma; HIF1α, hypoxia inducible factor-1α; NSCLC, non-small cell lung cancer; OC, ovarian cancer; OXPHOS, oxidative phosphorylation; ROS, Reactive Oxygen Species; SERPINB3, Serine Protease Inhibitor B3; TCF4, Transcription factor 4; TET3, Methyl cytosine dioxygenase 3; TMEM, transmembrane; TNBC, triple-negative breast cancer; YWHAZ, Tyrosine 3-monooxygenase/tryptophan 5-monooxygenase activation protein, zeta polypeptide.

Epigenetic modifications and DNA methylation

TMEM proteins are frequently regulated at the gene level by epigenetic modifications, such as DNA methylation and histone modifications (67-69). For example, TMEM7 and TMEM106A expression in HCC and GC is silenced by promoter methylation, which is associated with a poor prognosis (67-69,83,84). Studies have revealed that the downregulation of these TMEM proteins can be reversed through DNA demethylation, leading to the restoration of their tumor-suppressive functions (67-69,83,84). Similarly, TMEM25 in colon cancer and TMEM176A in esophageal cancer exhibit hypermethylation at their promoters, which suppresses their expression and promotes tumor progression (70,71,111,112). A common theme across these studies is that epigenetic changes, especially promoter methylation, not only regulate the expression and function of TMEM proteins but also influence treatment responses (112). For example, TMEM88 in OC has been associated with the differential sensitivity to chemotherapy based on its methylation status (223). Moreover, TMEM proteins such as TMEM173 and TMEM176B are involved in immune regulation, suggesting that reversing epigenetic silencing of TMEMs could offer dual therapeutic benefits: Restoring tumor suppressor activity and enhancing anti-tumor immune responses (97,100,101,119-121). Epigenetic silencing of TMEM proteins thus represents a notable mechanism of tumorigenesis and a promising target for therapeutic intervention.

Pathway level, EGFR and other signaling pathways

TMEM proteins are involved in the modulation of key signaling pathways that regulate cancer cell behaviors. For example, TMEM16A interacts with the EGFR in HNSCC and colon cancer, enhancing EGFR signaling and contributing to cancer cell proliferation, migration and survival (128). TMEM16A stabilizes EGFR and forms a positive feedback loop that potentiates cancer progression (128). Additionally, TMEM45A is associated with the TGF-β signaling pathway in OC, where it regulates proliferation and migration (154). TMEM158 has been implicated in modulating the Wnt/β-catenin signaling pathway in various types of cancer, facilitating tumor progression and metastasis (189). These interactions with key oncogenic pathways highlight the importance of TMEM proteins in driving cancer development and their potential as therapeutic targets.

Functional mechanisms, regulation of the cell cycle and apoptosis

TMEM proteins also directly influence cancer cell function, particularly in terms of regulating the cell cycle and apoptosis. For example, TMEM106A and TMEM158 have been shown to induce apoptosis in cancer cells by activating caspases and inactivating PARPs, which are involved in DNA damage repair (83,84,193,194). Additionally, TMEM45A regulates cell cycle progression and promotes cell proliferation by modulating the TGF-β and Notch pathways, which are central to tumor growth and invasion (157-160).

TME, hypoxia and immune regulation

TMEM proteins also interact with the TME, influencing cancer cell behavior and immune responses. TMEM173 (STING) carries out a key role in modulating immune responses within the TME by regulating innate immunity and immune surveillance (210-212). The activation of STING signaling enhances T cell infiltration into tumors, thereby potentiating antitumor immunity (210,212-214). However, in certain contexts, TMEM173 can contribute to immune suppression by promoting the differentiation of Tregs or inducing ER stress (213,216). TMEM158, another key protein, is involved in immune evasion by regulating immune checkpoints such as PD-1 and CTLA-4, thereby enhancing the ability of tumors to resist immune system attacks (190,197,198). Furthermore, TMEM proteins, such as TMEM205, might improve the prognosis of HCC patients by reducing the levels of immunosuppressive cells (M2 macrophages and Tregs) and facilitating the infiltration of cytotoxic T cells into the TME (207). These interactions with the TME contribute to tumor plasticity, immune evasion and resistance to chemotherapy. TMEM proteins influence cancer cell proliferation, migration and invasion through a combination of epigenetic regulation, modulation of key signaling pathways (such as EGFR, TGF-β and Wnt), functional mechanisms related to cell cycle regulation and apoptosis, and interactions with the TME. Their complex roles in cancer progression make them promising therapeutic targets, and further research into their molecular mechanisms could pave the way for novel cancer therapies.

The TMEM family of proteins carries out pivotal roles in mediating tumor chemoresistance through a range of mechanisms across multiple cancer types. TMEM45A is upregulated under hypoxic conditions and contributes to chemoresistance by inhibiting apoptosis and promoting EMT, with its effects varying by cancer type (156-161,221,222). TMEM158 enhances resistance by upregulating MDR proteins and suppressing apoptosis, particularly in colorectal and lung cancer (193,194). TMEM88 regulates the Wnt signaling pathway and is epigenetically modulated, with its overexpression associated with cisplatin resistance in OC (75,223). TMEM205 promotes cisplatin resistance via exosome-mediated drug efflux and by shaping the tumor immune microenvironment, particularly through macrophage polarization in GC (209,225-227). Collectively, these TMEM proteins are key regulators of drug resistance, making them promising targets for overcoming chemoresistance and improving the efficacy of chemotherapy. TMEM16A mediates platinum-based chemotherapy resistance in HNSCC by increasing oxidative stress and upregulating ATP7B, which facilitates lysosomal sequestration of platinum compounds and may serve as a biomarker to guide chemotherapy selection (229).

However, there are limitations to the current body of knowledge regarding the roles of TMEM proteins in cancer. Variations in study design, population demographics and methodological approaches introduce inconsistencies, leading to challenges in generalizing findings across different cancer types. Further research should address these limitations by employing standardized methods and exploring longitudinal analyses to improve the definition of the causal relationships between TMEM protein expression with tumor progression and chemoresistance. Future directions for research should focus on mechanistic studies to elucidate the molecular pathways through which TMEM proteins exert their effects on tumor biology and therapy resistance. With advancements in cryo-EM technology, researchers have overcome challenges such as low expression levels, hydrophobicity and conformational instability, enabling the high-resolution determination of several MP structures. Additionally, cryo-EM allows for the visualization of TPs at near-atomic resolution, providing detailed insights into their conformational states, ligand-binding sites and interactions with other proteins. This capability has greatly enhanced the understanding of TMEM protein structures and their roles in cellular signaling and chemotherapy resistance (20,230-232).

Chang et al (233) used cryo-EM to resolve the structure of a previously unsolved amyloid fibril composed of a 135-amino acid C-terminal fragment of TMEM106B, which is a common finding in various human neurodegenerative diseases. Cryo-EM allowed the determination of the TMEM106B fibril structure at a resolution of 2.7 Å from postmortem human brain tissue. The prevalence of amyloid fibrils made of TMEM106B across a range of debilitating human disorders suggests a shared fibrillization pathway that could initiate or accelerate neurodegeneration. Yuan et al (234) employed cryo-EM to analyze the structures of human GLUT4 bound to the small molecule inhibitor cytochalasin B, revealing an inward-open conformation at a resolution of 3.3 Å. Despite the nearly identical transmembrane domain conformation to GLUT1, the cryo-EM structure of GLUT4 reveals an extracellular glycosylation site and an intracellular helix that is invisible in the GLUT1 crystal structure. This structural study using cryo-EM lays the foundation for understanding the regulatory mechanisms of GLUT4 transport, and the methods developed for analyzing TMEMs will facilitate the structural determination of other small solute carrier. Shang et al (235) used cryo-EM to reveal the structural transitions of full-length STING/TMEM173 from both chicken and human. The structure was solved in its inactive dimeric state and in the dimeric and tetrameric states upon binding with cGAMP. These structural studies demonstrate how the transmembrane and cytoplasmic regions interact to form a domain-swapped dimeric assembly. Additionally, the closure of the ligand-binding domain, induced by cGAMP, triggers a conformational change in STING, ultimately facilitating the formation and activation of the STING oligomer. Zhang et al (236) reported four distinct cryo-EM structures revealing the apo states of IGF1R, a single transmembrane protein with a flexible structure. IGF1R carries out a key role in regulating cellular metabolism and growth. These conformations were classified as 'Resting states' and 'Active states', based on the orientation of α-CT helices and structural symmetry. Moreover, a 'Ligand-pocket' formed in the active conformations, providing a new perspective on the conformational changes of apo-IGF1R.

The structural insights gained from cryo-EM have considerable implications for drug screening and discovery, particularly in the identification of novel drug targets. For instance, cryo-EM structures of GPCRs and ion channels have unveiled detailed molecular interactions with drugs, enabling the rational design of therapeutics with improved specificity and efficacy (237-239). Cryo-EM has also facilitated the study of drug-target complexes, such as the interaction between statins and the liver transporter OATP1B1, offering a mechanistic understanding of drug transport and inhibition (240). High-resolution cryo-EM structures provide a solid foundation for structure-based drug design, enabling the rational development of small molecules, peptides and biologics that modulate membrane protein function. For example, the identification of ligand-binding sites and allosteric pockets in cryo-EM structures has guided the design of inhibitors and activators with enhanced specificity and potency (241,242). Moreover, cryo-EM studies of drug-resistant MPs, such as efflux transporters, have revealed the structural basis of resistance mechanisms. These insights can inform the development of next-generation therapeutics that overcome resistance by targeting alternative binding sites or modulating protein dynamics (243). Additionally, interventional studies focused on TMEM-targeted therapies, including epigenetic modulation and immune checkpoint inhibition, are essential to validate their therapeutic potential. Expanding on these findings with multi-omics approaches, such as genomics and proteomics, may uncover new interactions between TMEM proteins and the TME, thereby enhancing therapeutic efficacy.

Conclusions

The present review provides a timely and comprehensive synthesis of the roles that TMEM proteins carry out in cancer pathogenesis and therapy resistance, emphasizing their emerging potential as targets for diagnosis, prognosis and therapeutic intervention. Although individual TMEM family members have been studied, an integrative analysis of their structural features, physiological functions and regulatory mechanisms, particularly their dual roles in tumor progression and chemoresistance, has been lacking. The present review fills that gap through a systematic and multi-dimensional analysis across various types of cancer.

The involvement of TMEM proteins in tumorigenesis, the TME and drug resistance was discussed, with Table I and Fig. 4 offering summaries of their functional roles. Furthermore, the regulation of TMEM proteins via epigenetic modifications, their participation in key signaling pathways and their influence on cell cycle control, apoptosis, and immune regulation provide further insight into their multifaceted contributions to cancer biology.

Importantly, the present review highlights emerging TMEM proteins with clinical relevance and discuss how targeting these molecules could offer promising strategies to overcome therapeutic resistance, particularly in hypoxic tumors. These findings support the continued investigation of TMEM proteins as diagnostic biomarkers and therapeutic targets. Personalized treatment strategies based on TMEM expression profiles may improve therapeutic efficacy while minimizing adverse effects. Ultimately, advancing research on TMEM proteins could lead to innovative diagnostic tools and targeted therapies, improving clinical outcomes for patients with different types of cancer.

Availability of data and materials

Not applicable.

Authors' contributions

JS, DZ, BY, QL, HX and HP were responsible for the overall conceptualization and writing of the present review. JI, DZ, BY, QL, HX and HP contributed to drafting key content. HX and HP oversaw the critical review of key content and coordinated the review. All authors read and approved the final manuscript. Data authentication is not applicable.

Ethics approval and consent to participate

Not applicable.

Patient consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Abbreviations:

CC

cervical cancer

CRC

colorectal cancer

ESCC

esophageal squamous cell carcinoma

GBM

glioblastoma

GC

gastric cancer

HCC

hepatocellular carcinoma

MDR

multidrug resistance

MPs

membrane proteins

NSCLC

non-small cell lung cancer

OC

ovarian cancer

OS

overall survival

PC

pancreatic cancer

RCC

renal cell carcinoma

TC

thyroid cancer

TME

tumor microenvironment

TMEM

transmembrane

TNBC

triple-negative breast cancer

TPs

transmembrane proteins

HNSCC

head and neck squamous cell carcinoma

Acknowledgements

Not applicable.

Funding

This work was supported by the National Natural Science Foundation of China (grant no. 82173032,82203682), Liaoning Provincial Science and Technology Plan Project (grant no. 2023JH2/101700156), the Science and Technology Planning Project of Shenyang (grant no. 22-321-33-50), the Fundamental Research Funds for the Central Universities (LD202503).

References

1 

Scalise M and Jaakola VP: Membrane proteins: New approaches to probes, technologies, and drug design. SLAS Discov. 24:865–866. 2019. View Article : Google Scholar : PubMed/NCBI

2 

Curnow P: Designing minimalist membrane proteins. Biochem Soc Trans. 47:1233–1245. 2019. View Article : Google Scholar : PubMed/NCBI

3 

Pandey A, Shin K, Patterson RE, Liu XQ and Rainey JK: Current strategies for protein production and purification enabling membrane protein structural biology. Biochem Cell Biol. 94:507–527. 2016. View Article : Google Scholar : PubMed/NCBI

4 

Schmit K and Michiels C: TMEM proteins in cancer: A review. Front Pharmacol. 9:13452018. View Article : Google Scholar : PubMed/NCBI

5 

Marx S, Dal Maso T, Chen JW, Bury M, Wouters J, Michiels C and Le Calvé B: Transmembrane (TMEM) protein family members: Poorly characterized even if essential for the metastatic process. Semin Cancer Biol. 60:96–106. 2020. View Article : Google Scholar

6 

Akkafa F, Koyuncu İ, Temiz E, Dagli H, Dïlmec F and Akbas H: miRNA-mediated apoptosis activation through TMEM 48 inhibition in A549 cell line. Biochem Biophys Res Commun. 503:323–329. 2018. View Article : Google Scholar : PubMed/NCBI

7 

Sun Z, Zhao H, Yang S, Liu R, Yi L, Gao J, Liu S, Chen Y and Zhang Z: Edaravone Dexborneol protects against blood-brain barrier disruption following cerebral ischemia/reperfusion by upregulating pericyte coverage via vitronectin-integrin and PDGFB/PDGFR-β signaling. Free Radic Biol Med. 225:758–766. 2024. View Article : Google Scholar : PubMed/NCBI

8 

Sabbah DA, Hajjo R and Sweidan K: Review on epidermal growth factor receptor (EGFR) structure, signaling pathways, interactions, and recent updates of EGFR inhibitors. Curr Top Med Chem. 20:815–834. 2020. View Article : Google Scholar : PubMed/NCBI

9 

Krook MA, Reeser JW, Ernst G, Barker H, Wilberding M, Li G, Chen HZ and Roychowdhury S: Fibroblast growth factor receptors in cancer: Genetic alterations, diagnostics, therapeutic targets and mechanisms of resistance. Br J Cancer. 124:880–892. 2021. View Article : Google Scholar :

10 

Rodrigues Toledo C, Tantawy AA, Lima Fuscaldi L, Malavolta L and de Aguiar Ferreira C: EGFR- and integrin α(V)β(3)-targeting peptides as potential radiometal-labeled radiopharmaceuticals for cancer theranostics. Int J Mol Sci. 25:85532024. View Article : Google Scholar

11 

Chiaranunt P, Ferrara JL and Byersdorfer CA: Rethinking the paradigm: How comparative studies on fatty acid oxidation inform our understanding of T cell metabolism. Mol Immunol. 68C:564–574. 2015. View Article : Google Scholar

12 

Mohan CD, Rangappa KS and Sethi G: Transmembrane protein 25 abrogates monomeric EGFR-driven STAT3 activation in triple-negative breast cancer. MedComm (2020). 5:e4922024. View Article : Google Scholar : PubMed/NCBI

13 

Li S, Yue M, Xu H, Zhang X, Mao T, Quan M, Ma J, Wang Y, Ge W, Wang Y, et al: Chemotherapeutic drugs-induced pyroptosis mediated by gasdermin E promotes the progression and chemoresistance of pancreatic cancer. Cancer Lett. 564:2162062023. View Article : Google Scholar : PubMed/NCBI

14 

Lee W, Song G and Bae H: Review of synergistic anticancer effects of natural compounds combined with conventional therapeutics against chemoresistance and progression of pancreatic cancer. Mol Cell Toxicol. 21:455–471. 2025. View Article : Google Scholar

15 

Sato A, Takagi K, Yoshida M, Yamaguchi-Tanaka M, Sagehashi M, Miki Y, Miyashita M and Suzuki T: Discoidin domain receptor 2 contributes to breast cancer progression and chemoresistance by interacting with collagen type I. Cancers. 16:42852024. View Article : Google Scholar :

16 

Hegde RS and Keenan RJ: The mechanisms of integral membrane protein biogenesis. Nat Rev Mol Cell Biol. 23:107–124. 2022. View Article : Google Scholar

17 

Li F, Egea PF, Vecchio AJ, Asial I, Gupta M, Paulino J, Bajaj R, Dickinson MS, Ferguson-Miller S, Monk BC and Stroud RM: Highlighting membrane protein structure and function: A celebration of the Protein Data Bank. J Biol Chem. 296:1005572021. View Article : Google Scholar

18 

Tao X, Zhao C and MacKinnon R: Membrane protein isolation and structure determination in cell-derived membrane vesicles. Proc Natl Acad Sci USA. 120:e23023251202023. View Article : Google Scholar : PubMed/NCBI

19 

Vénien-Bryan C and Fernandes CAH: Overview of membrane protein sample preparation for single-particle cryo-electron microscopy analysis. Int J Mol Sci. 24:147852023. View Article : Google Scholar : PubMed/NCBI

20 

Thonghin N, Kargas V, Clews J and Ford RC: Cryo-electron microscopy of membrane proteins. Methods. 147:176–186. 2018. View Article : Google Scholar : PubMed/NCBI

21 

Gong J, Chen Y, Pu F, Sun P, He F, Zhang L, Li Y, Ma Z and Wang H: Understanding membrane protein drug targets in computational perspective. Curr Drug Targets. 20:551–564. 2019. View Article : Google Scholar

22 

Gulezian E, Crivello C, Bednenko J, Zafra C, Zhang Y, Colussi P and Hussain S: Membrane protein production and formulation for drug discovery. Trends Pharmacol Sci. 42:657–674. 2021. View Article : Google Scholar : PubMed/NCBI

23 

Dodd RB, Wilkinson T and Schofield DJ: Therapeutic monoclonal antibodies to complex membrane protein targets: Antigen generation and antibody discovery strategies. BioDrugs. 32:339–355. 2018. View Article : Google Scholar : PubMed/NCBI

24 

Gao S, Xu T, Wang W, Li J, Shan Y, Wang Y and Tan H: Polysaccharides from Anemarrhena asphodeloides Bge, the extraction, purification, structure characterization, biological activities and application of a traditional herbal medicine. Int J Biol Macromol. 311:1434972025. View Article : Google Scholar : PubMed/NCBI

25 

Gao S, Wang W, Li J, Wang Y, Shan Y and Tan H: Unveiling polysaccharides of Houttuynia cordata thunb: Extraction, purification, structure, bioactivities, and structure-activity relationships. Phytomedicine. 138:1564362025. View Article : Google Scholar

26 

Gao S, Li J, Wang W, Wang Y, Shan Y and Tan H: Rabdosia rubescens (Hemsl.) H. Hara: A potent anti-tumor herbal remedy - Botany, phytochemistry, and clinical applications and insights. J Ethnopharmacol. 340:1192002025. View Article : Google Scholar

27 

Chu XP and Xiong ZG: Acid-sensing ion channels in pathological conditions. Adv Exp Med Biol. 961:419–431. 2013. View Article : Google Scholar :

28 

Nandi PR: Pain in neurological conditions. Curr Opin Support Palliat Care. 6:194–200. 2012. View Article : Google Scholar : PubMed/NCBI

29 

Jones F, Gamper N and Gao H: Kv7 channels and excitability disorders. Handb Exp Pharmacol. 267:185–230. 2021. View Article : Google Scholar : PubMed/NCBI

30 

Wang X, Liu Y, Xue C, Hu Y, Zhao Y, Cai K, Li M and Luo Z: A protein-based cGAS-STING nanoagonist enhances T cell-mediated anti-tumor immune responses. Nat Commun. 13:56852022. View Article : Google Scholar : PubMed/NCBI

31 

Wang SP, Chen FY, Dong LX, Zhang YQ, Chen HY, Qiao K and Wang KJ: A novel innexin2 forming membrane hemichannel exhibits immune responses and cell apoptosis in Scylla paramamosain. Fish Shellfish Immunol. 47:485–499. 2015. View Article : Google Scholar : PubMed/NCBI

32 

Hauser AS, Attwood MM, Rask-Andersen M, Schiöth HB and Gloriam DE: Trends in GPCR drug discovery: New agents, targets and indications. Nat Rev Drug Discov. 16:829–842. 2017. View Article : Google Scholar : PubMed/NCBI

33 

Snijder HJ and Hakulinen J: Membrane protein production in E. coli for applications in drug discovery. Adv Exp Med Biol. 896:59–77. 2016. View Article : Google Scholar : PubMed/NCBI

34 

Kandasamy P, Gyimesi G, Kanai Y and Hediger MA: Amino acid transporters revisited: New views in health and disease. Trends Biochem Sci. 43:752–789. 2018. View Article : Google Scholar : PubMed/NCBI

35 

Kampen KR: Membrane proteins: the key players of a cancer cell. J Membr Biol. 242:69–74. 2011. View Article : Google Scholar : PubMed/NCBI

36 

Singh V, Kaur R, Kumari P, Pasricha C and Singh R: ICAM-1 and VCAM-1: Gatekeepers in various inflammatory and cardiovascular disorders. Clin Chim Acta. 548:1174872023. View Article : Google Scholar : PubMed/NCBI

37 

Pliego-Arreaga R, Cervantes-Montelongo JA, Silva-Martínez GA, Tristán-Flores FE, Pantoja-Hernández MA and Maldonado-Coronado JR: Joint hypermobility syndrome and membrane proteins: A comprehensive review. Biomolecules. 14:4722024. View Article : Google Scholar : PubMed/NCBI

38 

Guan Q, Bhowmick B, Upadhyay A and Han Q: Structure and functions of bacterial outer membrane protein A, a potential therapeutic target for bacterial infection. Curr Top Med Chem. 21:1129–1138. 2021. View Article : Google Scholar : PubMed/NCBI

39 

Kang H and Lee CJ: Transmembrane proteins with unknown function (TMEMs) as ion channels: Electrophysiological properties, structure, and pathophysiological roles. Exp Mol Med. 56:850–860. 2024. View Article : Google Scholar : PubMed/NCBI

40 

Li L and Li J: Dimerization of transmembrane proteins in cancer immunotherapy. Membranes. 13:3932023. View Article : Google Scholar : PubMed/NCBI

41 

Wesoly J, Pstrąg N, Derylo K, Michalec-Wawiórka B, Derebecka N, Nowicka H, Kajdasz A, Kluzek K, Srebniak M, Tchórzewski M, et al: Structural, topological, and functional characterization of transmembrane proteins TMEM213, 207, 116, 72 and 30B provides a potential link to ccRCC etiology. Am J Cancer Res. 13:1863–1883. 2023.PubMed/NCBI

42 

Nieto Gutierrez A and McDonald PH: GPCRs: Emerging anti-cancer drug targets. Cell Signal. 41:65–74. 2018. View Article : Google Scholar

43 

Dorsam RT and Gutkind JS: G-protein-coupled receptors and cancer. Nat Rev Cancer. 7:79–94. 2007. View Article : Google Scholar : PubMed/NCBI

44 

Lu X, Luo C, Wu J, Deng Y, Mu X, Zhang T, Yang X, Liu Q, Li Z, Tang S, et al: Ion channels and transporters regulate nutrient absorption in health and disease. J Cell Mol Med. 27:2631–2642. 2023. View Article : Google Scholar : PubMed/NCBI

45 

Talukdar S, Emdad L, Das SK and Fisher PB: EGFR: An essential receptor tyrosine kinase-regulator of cancer stem cells. Adv Cancer Res. 147:161–188. 2020. View Article : Google Scholar : PubMed/NCBI

46 

Liu Q, Huang J, Yan W, Liu Z, Liu S and Fang W: FGFR families: Biological functions and therapeutic interventions in tumors. MedComm (2020). 4:e3672023. View Article : Google Scholar : PubMed/NCBI

47 

Riordan R, Saxton A, McMillan PJ, Kow RL, Liachko NF and Kraemer BC: TMEM106B C-terminal fragments aggregate and drive neurodegenerative proteinopathy. bioRxiv [Preprint]: 2024.06.11.598478. 2024.

48 

Eber E, Trawinska-Bartnicka M, Sands D, Bellon G, Mellies U, Bolbás K, Quattrucci S, Mazurek H, Widmann R, Schoergenhofer C, et al: Aerosolized lancovutide in adolescents (≥12 years) and adults with cystic fibrosis-a randomized trial. J Cyst Fibros. 20:61–67. 2021. View Article : Google Scholar

49 

Vovdenko S, Morozov A, Ali S, Kogan E and Bezrukov E: Role of monocarboxylate transporters and glucose transporters in prostate cancer. Urologia. 90:491–498. 2023. View Article : Google Scholar

50 

Pérez-Herrero E and Fernández-Medarde A: Advanced targeted therapies in cancer: Drug nanocarriers, the future of chemotherapy. Eur J Pharm Biopharm. 93:52–79. 2015. View Article : Google Scholar : PubMed/NCBI

51 

Perneel J, Neumann M, Heeman B, Cheung S, Van den Broeck M, Wynants S, Baker M, Vicente CT, Faura J, Rademakers R and Mackenzie IRA: Accumulation of TMEM106B C-terminal fragments in neurodegenerative disease and aging. Acta Neuropathol. 145:285–302. 2023. View Article : Google Scholar

52 

Granholm AE, Englund E, Gilmore A, Head E, Yong WH, Perez SE, Guzman SJ, Hamlett ED and Mufson EJ: Neuropathological findings in Down syndrome, Alzheimer's disease and control patients with and without SARS-COV-2: Preliminary findings. Acta Neuropathol. 147:922024. View Article : Google Scholar : PubMed/NCBI

53 

Tran Q, Park J, Lee H, Hong Y, Hong S, Park S, Park J and Kim SH: TMEM39A and human diseases: A brief review. Toxicol Res. 33:205–209. 2017. View Article : Google Scholar : PubMed/NCBI

54 

Jankauskas SS, Varzideh F, Kansakar U, Al Tibi G, Densu Agyapong E, Gambardella J and Santulli G: Insights into molecular and cellular functions of the Golgi calcium/manganese-proton antiporter TMEM165. J Biol Chem. 300:1075672024. View Article : Google Scholar : PubMed/NCBI

55 

Dutta D, Kanca O, Shridharan RV, Marcogliese PC, Steger B, Morimoto M, Frost FG, Macnamara E; Undiagnosed Diseases Network; Wangler MF, et al: Loss of the endoplasmic reticulum protein Tmem208 affects cell polarity, development, and viability. Proc Natl Acad Sci USA. 121:e23225821212024. View Article : Google Scholar : PubMed/NCBI

56 

Hayez A, Roegiers E, Malaisse J, Balau B, Sterpin C, Achouri Y, De Rouvroit CL, Poumay Y, Michiels C and De Backer O: TMEM45A is dispensable for epidermal morphogenesis, keratinization and barrier formation. PLoS One. 11:e01470692016. View Article : Google Scholar : PubMed/NCBI

57 

Gallos G, Remy KE, Danielsson J, Funayama H, Fu XW, Chang HY, Yim P, Xu D and Emala CW Sr: Functional expression of the TMEM16 family of calcium-activated chloride channels in airway smooth muscle. Am J Physiol Lung Cell Mol Physiol. 305:L625–L634. 2013. View Article : Google Scholar : PubMed/NCBI

58 

Dulary E, Potelle S, Legrand D and Foulquier F: TMEM165 deficiencies in congenital disorders of glycosylation type II (CDG-II): Clues and evidences for roles of the protein in Golgi functions and ion homeostasis. Tissue Cell. 49A:150–156. 2017. View Article : Google Scholar

59 

Dodeller F, Gottar M, Huesken D, Iourgenko V and Cenni B: The lysosomal transmembrane protein 9B regulates the activity of inflammatory signaling pathways. J Biol Chem. 283:21487–21494. 2008. View Article : Google Scholar : PubMed/NCBI

60 

Herrera-Quiterio GA and Encarnación-Guevara S: The transmembrane proteins (TMEM) and their role in cell proliferation, migration, invasion, and epithelial-mesenchymal transition in cancer. Front Oncol. 13:12447402023. View Article : Google Scholar : PubMed/NCBI

61 

Jeon M, Yoo S, Park S, Choi Y, An J, Noh YR and Kim I: Over-expression of transmembrane protein 158 predicts aggressive tumor behavior and poor prognosis in lung cancer. Anticancer Res. 44:4885–4893. 2024. View Article : Google Scholar : PubMed/NCBI

62 

Pedemonte N and Galietta LJ: Structure and function of TMEM16 proteins (anoctamins). Physiol Rev. 94:419–459. 2014. View Article : Google Scholar : PubMed/NCBI

63 

Shi L, Wang X, Guo S, Gou H, Shang H, Jiang X, Wei C, Wang J, Li C, Wang L, et al: TMEM65 promotes gastric tumorigenesis by targeting YWHAZ to activate PI3K-Akt-mTOR pathway and is a therapeutic target. Oncogene. 43:931–943. 2024. View Article : Google Scholar : PubMed/NCBI

64 

Shepard HM, Phillips GL, D Thanos C and Feldmann M: Developments in therapy with monoclonal antibodies and related proteins. Clin Med. 17:220–232. 2017. View Article : Google Scholar

65 

Posner J, Barrington P, Brier T and Datta-Mannan A: Monoclonal antibodies: Past, present and future. Handb Exp Pharmacol. 260:81–141. 2019. View Article : Google Scholar : PubMed/NCBI

66 

Shaikh MH, Barrett JW, Khan MI, Kim HAJ, Zeng PYF, Mymryk JS and Nichols AC: Chromosome 3p loss in the progression and prognosis of head and neck cancer. Oral Oncol. 109:1049442020. View Article : Google Scholar : PubMed/NCBI

67 

Zhou X, Popescu NC, Klein G and Imreh S: The interferon-alpha responsive gene TMEM7 suppresses cell proliferation and is downregulated in human hepatocellular carcinoma. Cancer Genet Cytogenet. 177:6–15. 2007. View Article : Google Scholar : PubMed/NCBI

68 

Kholodnyuk ID, Kozireva S, Kost-Alimova M, Kashuba V, Klein G and Imreh S: Down regulation of 3p genes, LTF, SLC38A3 and DRR1, upon growth of human chromosome 3-mouse fibrosarcoma hybrids in severe combined immunodeficiency mice. Int J Cancer. 119:99–107. 2006. View Article : Google Scholar : PubMed/NCBI

69 

Kiss H, Darai E, Kiss C, Kost-Alimova M, Klein G, Dumanski JP and Imreh S: Comparative human/murine sequence analysis of the common eliminated region 1 from human 3p21.3. Mamm Genome. 13:646–655. 2002. View Article : Google Scholar : PubMed/NCBI

70 

Hrašovec S, Hauptman N, Glavač D, Jelenc F and Ravnik-Glavač M: TMEM25 is a candidate biomarker methylated and down-regulated in colorectal cancer. Dis Markers. 34:93–104. 2013. View Article : Google Scholar

71 

Bi J, Wu Z, Zhang X, Zeng T, Dai W, Qiu N, Xu M, Qiao Y, Ke L, Zhao J, et al: TMEM25 inhibits monomeric EGFR-mediated STAT3 activation in basal state to suppress triple-negative breast cancer progression. Nat Commun. 14:23422023. View Article : Google Scholar : PubMed/NCBI

72 

Doolan P, Clynes M, Kennedy S, Mehta JP, Germano S, Ehrhardt C, Crown J and O'Driscoll L: TMEM25, REPS2 and Meis 1: Favourable prognostic and predictive biomarkers for breast cancer. Tumour Biol. 30:200–209. 2009. View Article : Google Scholar : PubMed/NCBI

73 

Xi P, Zhang Z, Liu Y, Nie Y, Gong B, Liu J, Huang H, Liu Z, Sun T and Xie W: Multidimensional comprehensive and integrated analysis of the potential function of TMEM25 in renal clear cell carcinoma with low expression status. Aging (Albany NY). 16:367–388. 2024.PubMed/NCBI

74 

Cai M, Ni WJ, Wang YH, Wang JJ and Zhou H: Targeting TMEM88 as an attractive therapeutic strategy in malignant tumors. Front Oncol. 12:9063722022. View Article : Google Scholar : PubMed/NCBI

75 

Ge YX, Wang CH, Hu FY, Pan LX, Min J, Niu KY, Zhang L, Li J and Xu T: New advances of TMEM88 in cancer initiation and progression, with special emphasis on Wnt signaling pathway. J Cell Physiol. 233:79–87. 2018. View Article : Google Scholar

76 

Li LY, Yang CC, Li SW, Liu YM, Li HD, Hu S, Zhou H, Wang JL, Shen H, Meng XM, et al: TMEM88 modulates the secretion of inflammatory factors by regulating YAP signaling pathway in alcoholic liver disease. Inflamm Res. 69:789–800. 2020. View Article : Google Scholar : PubMed/NCBI

77 

Geng Q, Chen X and Chen N: Transmembrane protein 88 exerts a tumor-inhibitory role in thyroid cancer through restriction of Wnt/β-catenin signaling. Exp Cell Res. 395:1121932020. View Article : Google Scholar

78 

Nusse R and Clevers H: Wnt/β-Catenin signaling, disease, and emerging therapeutic modalities. Cell. 169:985–999. 2017. View Article : Google Scholar : PubMed/NCBI

79 

Zhao X, Li G, Chong T, Xue L, Luo Q, Tang X, Zhai X, Chen J and Zhang X: TMEM88 exhibits an antiproliferative and anti-invasive effect in bladder cancer by downregulating Wnt/β-catenin signaling. J Biochem Mol Toxicol. 35:e228352021. View Article : Google Scholar

80 

Zhang X, Yu X, Jiang G, Miao Y, Wang L, Zhang Y, Liu Y, Fan C, Lin X, Dong Q, et al: Cytosolic TMEM88 promotes invasion and metastasis in lung cancer cells by binding DVLS. Cancer Res. 75:4527–4537. 2015. View Article : Google Scholar : PubMed/NCBI

81 

Ma R, Feng N, Yu X, Lin H, Zhang X, Shi O, Zhang H, Zhang S, Li L, Zheng M, et al: Promoter methylation of Wnt/β-Catenin signal inhibitor TMEM88 is associated with unfavorable prognosis of non-small cell lung cancer. Cancer Biol Med. 14:377–386. 2017. View Article : Google Scholar

82 

Yu X, Zhang X, Zhang Y, Jiang G, Mao X and Jin F: Cytosolic TMEM88 promotes triple-negative breast cancer by interacting with Dvl. Oncotarget. 6:25034–25045. 2015. View Article : Google Scholar : PubMed/NCBI

83 

Mao D, Yan F, Zhang X and Gao G: TMEM106A inhibits enveloped virus release from cell surface. iScience. 25:1038432022. View Article : Google Scholar : PubMed/NCBI

84 

Xu D, Qu L, Hu J, Li G, Lv P, Ma D, Guo M and Chen Y: Transmembrane protein 106A is silenced by promoter region hypermethylation and suppresses gastric cancer growth by inducing apoptosis. J Cell Mol Med. 18:1655–1666. 2014. View Article : Google Scholar : PubMed/NCBI

85 

Choi B, Han TS, Min J, Hur K, Lee SM, Lee HJ, Kim YJ and Yang HK: MAL and TMEM220 are novel DNA methylation markers in human gastric cancer. Biomarkers. 22:35–44. 2017. View Article : Google Scholar

86 

Wu C, Xu J, Wang H, Zhang J, Zhong J, Zou X, Chen Y, Yang G, Zhong Y, Lai D, et al: TMEM106a is a novel tumor suppressor in human renal cancer. Kidney Blood Press Res. 42:853–864. 2017. View Article : Google Scholar : PubMed/NCBI

87 

Shi S, Wang B, Wan J, Song L, Zhu G, Du J, Ye L, Zhao Q, Cai J, Chen Q, et al: TMEM106A transcriptionally regulated by promoter methylation is involved in invasion and metastasis of hepatocellular carcinoma. Acta Biochim Biophys Sin. 54:1008–1020. 2022. View Article : Google Scholar : PubMed/NCBI

88 

Liu J and Zhu H: TMEM106A inhibits cell proliferation, migration, and induces apoptosis of lung cancer cells. J Cell Biochem. 120:7825–7833. 2019. View Article : Google Scholar

89 

Hou J, Niu Y, Yan J, Tian J, Yu W, Zhang G, Li T and Wang Z: Non-invasive diagnosis for urothelial carcinoma using a dual-target DNA methylation biomarker panel. Clin Chim Acta. 569:1201642025. View Article : Google Scholar : PubMed/NCBI

90 

Guo X, Zeng S, Ji X, Meng X, Lei N, Yang H and Mu X: Type I Interferon-induced TMEM106A blocks attachment of EV-A71 virus by interacting with the membrane protein SCARB2. Front Immunol. 13:8178352022. View Article : Google Scholar : PubMed/NCBI

91 

Zhang X, Feng T, Zhou X, Sullivan PM, Hu F, Lou Y, Yu J, Feng J, Liu H and Chen Y: Inactivation of TMEM106A promotes lipopolysaccharide-induced inflammation via the MAPK and NF-κB signaling pathways in macrophages. Clin Exp Immunol. 203:125–136. 2021. View Article : Google Scholar

92 

Sun Z and Hornung V: cGAS-STING signaling. Curr Biol. 32:R730–R734. 2022. View Article : Google Scholar : PubMed/NCBI

93 

Chen C and Xu P: Cellular functions of cGAS-STING signaling. Trends Cell Biol. 33:630–648. 2023. View Article : Google Scholar

94 

Luo X, Li H, Ma L, Zhou J, Guo X, Woo SL, Pei Y, Knight LR, Deveau M, Chen Y, et al: Expression of STING is increased in liver tissues from patients with NAFLD and promotes macrophage-mediated hepatic inflammation and fibrosis in mice. Gastroenterology. 155:1971–1984.e4. 2018. View Article : Google Scholar : PubMed/NCBI

95 

Yu Y, Liu Y, An W, Song J, Zhang Y and Zhao X: STING-mediated inflammation in Kupffer cells contributes to progression of nonalcoholic steatohepatitis. J Clin Invest. 129:546–555. 2019. View Article : Google Scholar :

96 

Chin KH, Tu ZL, Su YC, Yu YJ, Chen HC, Lo YC, Chen CP, Barber GN, Chuah ML, Liang ZX and Chou SH: Novel c-di-GMP recognition modes of the mouse innate immune adaptor protein STING. Acta Crystallogr D Biol Crystallogr. 69:352–366. 2013. View Article : Google Scholar : PubMed/NCBI

97 

Liu D, Liang S, Ma K, Meng QF, Li X, Wei J, Zhou M, Yun K, Pan Y, Rao L, et al: Tumor microenvironment-responsive nanoparticles amplifying STING signaling pathway for cancer immunotherapy. Adv Mater. 36:e23048452024. View Article : Google Scholar

98 

Garland KM, Sheehy TL and Wilson JT: Chemical and biomolecular strategies for STING pathway activation in cancer immunotherapy. Chem Rev. 122:5977–6039. 2022. View Article : Google Scholar : PubMed/NCBI

99 

Liang S, Ma H, Liu Y, Hai L, Tian Y, Sun Y and Wang Z: Nano-immunomodulator amplifies STING activation in tumor-associated macrophages for cancer immunotherapy. J Control Release. 383:1138462025. View Article : Google Scholar : PubMed/NCBI

100 

Deng A, Fan R, Hai Y, Zhuang J, Zhang B, Lu X, Wang W, Luo L, Bai G, Liang L, et al: A STING agonist prodrug reprograms tumor-associated macrophage to boost colorectal cancer immunotherapy. Theranostics. 15:277–299. 2025. View Article : Google Scholar : PubMed/NCBI

101 

Ishikawa T, Tamura E, Kasahara M, Uchida H, Higuchi M, Kobayashi H, Shimizu H, Ogawa E, Yotani N, Irie R, et al: Severe liver disorder following liver transplantation in STING-Associated vasculopathy with onset in infancy. J Clin Immunol. 41:967–974. 2021. View Article : Google Scholar : PubMed/NCBI

102 

Song H, Chen L, Pan X, Shen Y, Ye M, Wang G, Cui C, Zhou Q, Tseng Y, Gong Z, et al: Targeting tumor monocyte-intrinsic PD-L1 by rewiring STING signaling and enhancing STING agonist therapy. Cancer Cell. 43:503–518.e0. 2025. View Article : Google Scholar : PubMed/NCBI

103 

Li S, Mirlekar B, Johnson BM, Brickey WJ, Wrobel JA, Yang N, Song D, Entwistle S, Tan X, Deng M, et al: STING-induced regulatory B cells compromise NK function in cancer immunity. Nature. 610:373–380. 2022. View Article : Google Scholar : PubMed/NCBI

104 

Larabi A, Barnich N and Nguyen HTT: New insights into the interplay between autophagy, gut microbiota and inflammatory responses in IBD. Autophagy. 16:38–51. 2020. View Article : Google Scholar :

105 

Bu Y, Liu F, Jia QA and Yu SN: Decreased expression of TMEM173 predicts poor prognosis in patients with hepatocellular carcinoma. PLoS One. 11:e01656812016. View Article : Google Scholar : PubMed/NCBI

106 

Wang J, Li S, Wang M, Wang X, Chen S, Sun Z, Ren X, Huang G, Sumer BD, Yan N, et al: STING licensing of type I dendritic cells potentiates antitumor immunity. Sci Immunol. 9:eadj39452024. View Article : Google Scholar : PubMed/NCBI

107 

Lai P, Liu L, Bancaro N, Troiani M, Calì B, Li Y, Chen J, Singh PK, Arzola RA, Attanasio G, et al: Mitochondrial DNA released by senescent tumor cells enhances PMN-MDSC-driven immunosuppression through the cGAS-STING pathway. Immunity. 58:811–825.e7. 2025. View Article : Google Scholar : PubMed/NCBI

108 

Zeng Q, Liu M, Wang Z, Zhou R and Ai K: Enhancing radiotherapy-induced anti-tumor immunity via nanoparticle-mediated STING agonist synergy. Mol Cancer. 24:1762025. View Article : Google Scholar : PubMed/NCBI

109 

Riegman PH, Burgart LJ, Wang KK, Wink-Godschalk JC, Dinjens WN, Siersema PD, Tilanus HW and van Dekken H: Allelic imbalance of 7q32.3-q36.1 during tumorigenesis in Barrett's esophagus. Cancer Res. 62:1531–1533. 2002.PubMed/NCBI

110 

Hill M, Russo S, Olivera D, Malcuori M, Galliussi G and Segovia M: The intracellular cation channel TMEM176B as a dual immunoregulator. Front Cell Dev Biol. 10:10384292022. View Article : Google Scholar : PubMed/NCBI

111 

Wang Y, Zhang Y, Herman JG, Linghu E and Guo M: Epigenetic silencing of TMEM176A promotes esophageal squamous cell cancer development. Oncotarget. 8:70035–70048. 2017. View Article : Google Scholar : PubMed/NCBI

112 

Gao D, Han Y, Yang Y, Herman JG, Linghu E, Zhan Q, Fuks F, Lu ZJ and Guo M: Methylation of TMEM176A is an independent prognostic marker and is involved in human colorectal cancer development. Epigenetics. 12:575–583. 2017. View Article : Google Scholar : PubMed/NCBI

113 

Liu Z, An H, Song P, Wang D, Li S, Chen K and Pang Q: Potential targets of TMEM176A in the growth of glioblastoma cells. Onco Targets Ther. 11:7763–7775. 2018. View Article : Google Scholar : PubMed/NCBI

114 

Qian W, Xu CY, Hong W, Li ZM and Xu DG: Transmembrane protein 176B promotes epithelial-mesenchymal transition in colorectal cancer through inflammasome inhibition. World J Gastrointest Oncol. 17:976732025. View Article : Google Scholar : PubMed/NCBI

115 

Li J, Fang Z, Dal E, Zhang H, Yu K, Ma M, Wang M, Sun R, Lu M, Wang H and Li Y: Transmembrane protein 176B regulates amino acid metabolism through the PI3K-Akt-mTOR signaling pathway and promotes gastric cancer progression. Cancer Cell Int. 24:952024. View Article : Google Scholar : PubMed/NCBI

116 

Yan L, Song Z, Yi L, Tian C, Zhang R, Qin X, Wang X, Ren S, Ma X, Wang X, et al: TMEM176B inhibits ovarian cancer progression by regulating EMT via the Wnt/β-catenin signaling pathway. J Transl Med. 23:3502025. View Article : Google Scholar

117 

Condamine T, Le Texier L, Howie D, Lavault A, Hill M, Halary F, Cobbold S, Waldmann H, Cuturi MC and Chiffoleau E: Tmem176B and Tmem176A are associated with the immature state of dendritic cells. J Leukoc Biol. 88:507–515. 2010. View Article : Google Scholar : PubMed/NCBI

118 

Segovia M, Louvet C, Charnet P, Savina A, Tilly G, Gautreau L, Carretero-Iglesia L, Beriou G, Cebrian I, Cens T, et al: Autologous dendritic cells prolong allograft survival through Tmem176b-dependent antigen cross-presentation. Am J Transplant. 14:1021–1031. 2014. View Article : Google Scholar : PubMed/NCBI

119 

Segovia M, Russo S, Jeldres M, Mahmoud YD, Perez V, Duhalde M, Charnet P, Rousset M, Victoria S, Veigas F, et al: Targeting TMEM176B enhances antitumor immunity and augments the efficacy of immune checkpoint blockers by unleashing inflammasome activation. Cancer Cell. 35:767–781.e6. 2019. View Article : Google Scholar : PubMed/NCBI

120 

Victoria S, Castro A, Pittini A, Olivera D, Russo S, Cebrian I, Mombru AW, Osinaga E, Pardo H, Segovia M and Hill M: Formulating a TMEM176B blocker in chitosan nanoparticles uncouples its paradoxical roles in innate and adaptive antitumoral immunity. Int J Biol Macromol. 279:1353272024. View Article : Google Scholar : PubMed/NCBI

121 

Jiang L, Yang Y, Liu F, Ma M, Gao J, Sun L, Chen Y, Shen Z and Wu D: A potential diagnostic and prognostic biomarker TMEM176B and its relationship with immune infiltration in skin cutaneous melanoma. Front Cell Dev Biol. 10:8599582022. View Article : Google Scholar : PubMed/NCBI

122 

Jing L, An Y, Cai T, Xiang J, Li B, Guo J, Ma X, Wei L, Tian Y, Cheng X, et al: A subpopulation of CD146(+) macrophages enhances antitumor immunity by activating the NLRP3 inflammasome. Cell Mol Immunol. 20:908–923. 2023. View Article : Google Scholar : PubMed/NCBI

123 

Picotto G, Morse LR, Nguyen N, Saltzman J and Battaglino R: TMEM176A and TMEM176B are candidate regulators of inhibition of dendritic cell maturation and function after chronic spinal cord injury. J Neurotrauma. 37:528–533. 2020. View Article : Google Scholar :

124 

Cuajungco MP, Podevin W, Valluri VK, Bui Q, Nguyen VH and Taylor K: Abnormal accumulation of human transmembrane (TMEM)-176A and 176B proteins is associated with cancer pathology. Acta Histochem. 114:705–712. 2012. View Article : Google Scholar : PubMed/NCBI

125 

Li H, Zhang M, Linghu E, Zhou F, Herman JG, Hu L and Guo M: Epigenetic silencing of TMEM176A activates ERK signaling in human hepatocellular carcinoma. Clin Epigenetics. 10:1372018. View Article : Google Scholar : PubMed/NCBI

126 

Al-Hosni R, Ilkan Z, Agostinelli E and Tammaro P: The pharmacology of the TMEM16A channel: therapeutic opportunities. Trends Pharmacol Sci. 43:712–725. 2022. View Article : Google Scholar : PubMed/NCBI

127 

Okuyama K and Yanamoto S: TMEM16A as a potential treatment target for head and neck cancer. J Exp Clin Cancer Res. 41:1962022. View Article : Google Scholar : PubMed/NCBI

128 

Bill A, Gutierrez A, Kulkarni S, Kemp C, Bonenfant D, Voshol H, Duvvuri U and Gaither LA: ANO1/TMEM16A interacts with EGFR and correlates with sensitivity to EGFR-targeting therapy in head and neck cancer. Oncotarget. 6:9173–9188. 2015. View Article : Google Scholar : PubMed/NCBI

129 

Wang F, Zhang Y, Gao M and Zeng X: TMEM16A inhibits renal tubulointerstitial fibrosis via Wnt/β-catenin signaling during hypertension nephropathy. Cell Signal. 117:1110882024. View Article : Google Scholar

130 

Yan Y, Ding X, Han C, Gao J, Liu Z, Liu Y and Wang K: Involvement of TMEM16A/ANO1 upregulation in the oncogenesis of colorectal cancer. Biochim Biophys Acta Mol Basis Dis. 1868:1663702022. View Article : Google Scholar : PubMed/NCBI

131 

Li H, Yu Z, Wang H, Wang N, Sun X, Yang S, Hua X and Liu Z: Role of ANO1 in tumors and tumor immunity. J Cancer Res Clin Oncol. 148:2045–2068. 2022. View Article : Google Scholar : PubMed/NCBI

132 

Filippou A, Pehkonen H, Karhemo PR, Väänänen J, Nieminen AI, Klefström J, Grénman R, Mäkitie AA, Joensuu H and Monni O: ANO1 expression orchestrates p27Kip1/MCL1-mediated signaling in head and neck squamous cell carcinoma. Cancers. 13:11702021. View Article : Google Scholar : PubMed/NCBI

133 

Shiwarski DJ, Shao C, Bill A, Kim J, Xiao D, Bertrand CA, Seethala RS, Sano D, Myers JN, Ha P, et al: To 'grow' or 'go': TMEM16A expression as a switch between tumor growth and metastasis in SCCHN. Clin Cancer Res. 20:4673–4688. 2014. View Article : Google Scholar : PubMed/NCBI

134 

Godse NR, Khan N, Yochum ZA, Gomez-Casal R, Kemp C, Shiwarski DJ, Seethala RS, Kulich S, Seshadri M, Burns TF and Duvvuri U: TMEM16A/ANO1 inhibits apoptosis via downregulation of bim expression. Clin Cancer Res. 23:7324–7332. 2017. View Article : Google Scholar : PubMed/NCBI

135 

Ayoub C, Wasylyk C, Li Y, Thomas E, Marisa L, Robé A, Roux M, Abecassis J, de Reyniès A and Wasylyk B: ANO1 amplification and expression in HNSCC with a high propensity for future distant metastasis and its functions in HNSCC cell lines. Br J Cancer. 103:715–726. 2010. View Article : Google Scholar : PubMed/NCBI

136 

Dixit R, Kemp C, Kulich S, Seethala R, Chiosea S, Ling S, Ha PK and Duvvuri U: TMEM16A/ANO1 is differentially expressed in HPV-negative versus HPV-positive head and neck squamous cell carcinoma through promoter methylation. Sci Rep. 5:166572015. View Article : Google Scholar : PubMed/NCBI

137 

Liu F, Cao QH, Lu DJ, Luo B, Lu XF, Luo RC and Wang XG: TMEM16A overexpression contributes to tumor invasion and poor prognosis of human gastric cancer through TGF-β signaling. Oncotarget. 6:11585–11599. 2015. View Article : Google Scholar : PubMed/NCBI

138 

Britschgi A, Bill A, Brinkhaus H, Rothwell C, Clay I, Duss S, Rebhan M, Raman P, Guy CT, Wetzel K, et al: Calcium-activated chloride channel ANO1 promotes breast cancer progression by activating EGFR and CAMK signaling. Proc Natl Acad Sci USA. 110:E1026–E1034. 2013. View Article : Google Scholar : PubMed/NCBI

139 

Bae JS, Park JY, Park SH, Ha SH, An AR, Noh SJ, Kwon KS, Jung SH, Park HS, Kang MJ and Jang KY: Expression of ANO1/DOG1 is associated with shorter survival and progression of breast carcinomas. Oncotarget. 9:607–621. 2018. View Article : Google Scholar : PubMed/NCBI

140 

Rhodes A, Jasani B, Balaton AJ, Barnes DM and Miller KD: Frequency of oestrogen and progesterone receptor positivity by immunohistochemical analysis in 7016 breast carcinomas: correlation with patient age, assay sensitivity, threshold value, and mammographic screening. J Clin Pathol. 53:688–696. 2000. View Article : Google Scholar : PubMed/NCBI

141 

Meng Y, Liu Z, Zhai C, Di T, Zhang L, Zhang L, Xie X, Lin Y, Wang N, Zhao J, et al: Paeonol inhibits the development of 1-chloro-2,4-dinitrobenzene-induced atopic dermatitis via mast and T cells in BALB/c mice. Mol Med Rep. 19:3217–3229. 2019.PubMed/NCBI

142 

Sui Y, Sun M, Wu F, Yang L, Di W, Zhang G, Zhong L, Ma Z, Zheng J, Fang X and Ma T: Inhibition of TMEM16A expression suppresses growth and invasion in human colorectal cancer cells. PLoS One. 9:e1154432014. View Article : Google Scholar : PubMed/NCBI

143 

Li H, Yang Q, Huo S, Du Z, Wu F, Zhao H, Chen S, Yang L, Ma Z and Sui Y: Expression of TMEM16A in colorectal cancer and its correlation with clinical and pathological parameters. Front Oncol. 11:6522622021. View Article : Google Scholar : PubMed/NCBI

144 

Liu W, Lu M, Liu B, Huang Y and Wang K: Inhibition of Ca(2+)-activated Cl(-) channel ANO1/TMEM16A expression suppresses tumor growth and invasiveness in human prostate carcinoma. Cancer Lett. 326:41–51. 2012. View Article : Google Scholar : PubMed/NCBI

145 

Song Y, Gao J, Guan L, Chen X, Gao J and Wang K: Inhibition of ANO1/TMEM16A induces apoptosis in human prostate carcinoma cells by activating TNF-α signaling. Cell Death Dis. 9:7032018. View Article : Google Scholar

146 

Jeon D, Jo M, Lee Y, Park SH, Phan HTL, Nam JH and Namkung W: Inhibition of ANO1 by cis- and trans-resveratrol and their anticancer activity in human prostate cancer PC-3 cells. Int J Mol Sci. 24:11862023. View Article : Google Scholar : PubMed/NCBI

147 

Seo Y, Ryu K, Park J, Jeon DK, Jo S, Lee HK and Namkung W: Inhibition of ANO1 by luteolin and its cytotoxicity in human prostate cancer PC-3 cells. PLoS One. 12:e01749352017. View Article : Google Scholar : PubMed/NCBI

148 

Liu Z, Zhang S, Hou F, Zhang C, Gao J and Wang K: Inhibition of Ca(2+) -activated chloride channel ANO1 suppresses ovarian cancer through inactivating PI3K/Akt signaling. Int J Cancer. 144:2215–2226. 2019. View Article : Google Scholar

149 

Bai X, Cheng Y, Wan H, Li S, Kang X and Guo S: Natural compound allicin containing thiosulfinate moieties as transmembrane protein 16A (TMEM16A) Ion channel inhibitor for food adjuvant therapy of lung cancer. J Agric Food Chem. 71:535–545. 2023. View Article : Google Scholar

150 

Seo Y, Park J, Kim M, Lee HK, Kim JH, Jeong JH and Namkung W: Inhibition of ANO1/TMEM16A chloride channel by idebenone and its cytotoxicity to cancer cell lines. PLoS One. 10:e01336562015. View Article : Google Scholar : PubMed/NCBI

151 

Yin L, Menon R, Gupta R, Vaught L, Okunieff P and Vidyasagar S: Glucose enhances rotavirus enterotoxin-induced intestinal chloride secretion. Pflugers Arch. 469:1093–1105. 2017. View Article : Google Scholar : PubMed/NCBI

152 

Hayez A, Malaisse J, Roegiers E, Reynier M, Renard C, Haftek M, Geenen V, Serre G, Simon M, de Rouvroit CL, et al: High TMEM45A expression is correlated to epidermal keratinization. Exp Dermatol. 23:339–344. 2014. View Article : Google Scholar : PubMed/NCBI

153 

Sun W, Qiu G, Zou Y, Cai Z, Wang P, Lin X, Huang J, Jiang L, Ding X and Hu G: Knockdown of TMEM45A inhibits the proliferation, migration and invasion of glioma cells. Int J Clin Exp Pathol. 8:12657–12667. 2015.

154 

Guo J, Chen L, Luo N, Yang W, Qu X and Cheng Z: Inhibition of TMEM45A suppresses proliferation, induces cell cycle arrest and reduces cell invasion in human ovarian cancer cells. Oncol Rep. 33:3124–3130. 2015. View Article : Google Scholar : PubMed/NCBI

155 

Manawapat-Klopfer A, Thomsen LT, Martus P, Munk C, Russ R, Gmuender H, Frederiksen K, Haedicke-Jarboui J, Stubenrauch F, Kjaer SK and Iftner T: TMEM45A, SERPINB5 and p16INK4A transcript levels are predictive for development of high-grade cervical lesions. Am J Cancer Res. 6:1524–1536. 2016.PubMed/NCBI

156 

Liu Y, Liu L and Mou ZX: TMEM45A affects proliferation, apoptosis, epithelial-mesenchymal transition, migration, invasion and cisplatin resistance of HPV-Positive cervical cancer cell lines. Biochem Genet. 60:173–190. 2022. View Article : Google Scholar

157 

Neuperger P, Balog J, Tiszlavicz L, Furák J, Gémes N, Kotogány E, Szalontai K, Puskás LG and Szebeni GJ: Analysis of the single-cell heterogeneity of adenocarcinoma cell lines and the investigation of intratumor heterogeneity reveals the expression of transmembrane protein 45A (TMEM45A) in lung adenocarcinoma cancer patients. Cancers. 14:1442021. View Article : Google Scholar

158 

Jiang H, Chen H, Wan P, Liang M and Chen N: Upregulation of TMEM45A promoted the progression of clear cell renal cell carcinoma in vitro. J Inflamm Res. 14:6421–6430. 2021. View Article : Google Scholar : PubMed/NCBI

159 

Xie Q, Guo T, Deng H, Yu C and Fang C: Pan-cancer analysis of TMEM45A and exploration of its prognostic value and mechanism in gastric cancer. Cell Mol Biol. 70:218–229. 2024. View Article : Google Scholar : PubMed/NCBI

160 

Chen C, Chen Z, Zhao J, Wen X, Yao H, Weng Z, Xiong H, Zheng Z and Wu J: TMEM45A enhances palbociclib resistance and cellular glycolysis by activating AKT/mTOR signaling pathway in HR+ breast cancer. Cell Death Discov. 11:472025. View Article : Google Scholar

161 

Flamant L, Roegiers E, Pierre M, Hayez A, Sterpin C, De Backer O, Arnould T, Poumay Y and Michiels C: TMEM45A is essential for hypoxia-induced chemoresistance in breast and liver cancer cells. BMC Cancer. 12:3912012. View Article : Google Scholar : PubMed/NCBI

162 

Zhu M, Jiang B, Yan D, Wang X, Ge H and Sun Y: Knockdown of TMEM45A overcomes multidrug resistance and epithelial-mesenchymal transition in human colorectal cancer cells through inhibition of TGF-β signalling pathway. Clin Exp Pharmacol Physiol. 47:503–516. 2020. View Article : Google Scholar

163 

Okada N, Yamamoto T, Watanabe M, Yoshimura Y, Obana E, Yamazaki N, Kawazoe K, Shinohara Y and Minakuchi K: Identification of TMEM45B as a protein clearly showing thermal aggregation in SDS-PAGE gels and dissection of its amino acid sequence responsible for this aggregation. Protein Expr Purif. 77:118–123. 2011. View Article : Google Scholar : PubMed/NCBI

164 

Hu R, Hu F, Xie X, Wang L, Li G, Qiao T, Wang M and Xiao H: TMEM45B, up-regulated in human lung cancer, enhances tumorigenicity of lung cancer cells. Tumour Biol. 37:12181–12191. 2016. View Article : Google Scholar : PubMed/NCBI

165 

Zhao LC, Shen BY, Deng XX, Chen H, Zhu ZG and Peng CH: TMEM45B promotes proliferation, invasion and migration and inhibits apoptosis in pancreatic cancer cells. Mol Biosyst. 12:1860–1870. 2016. View Article : Google Scholar : PubMed/NCBI

166 

Shen K, Yu W, Yu Y, Liu X and Cui X: Knockdown of TMEM45B inhibits cell proliferation and invasion in gastric cancer. Biomed Pharmacother. 104:576–581. 2018. View Article : Google Scholar : PubMed/NCBI

167 

Li Y, Guo W, Liu S, Zhang B, Yu BB, Yang B, Kan SL and Feng SQ: Silencing transmembrane protein 45B (TNEM45B) inhibits proliferation, invasion, and tumorigenesis in osteosarcoma cells. Oncol Res. 25:1021–1026. 2017. View Article : Google Scholar : PubMed/NCBI

168 

Tanioku T, Nishibata M, Tokinaga Y, Konno K, Watanabe M, Hemmi H, Fukuda-Ohta Y, Kaisho T, Furue H and Kawamata T: Tmem45b is essential for inflammation- and tissue injury-induced mechanical pain hypersensitivity. Proc Natl Acad Sci USA. 119:e21219891192022. View Article : Google Scholar : PubMed/NCBI

169 

Luo F, Yang K, Wang YZ and Lin D: TMEM45B is a novel predictive biomarker for prostate cancer progression and metastasis. Neoplasma. 65:815–821. 2018. View Article : Google Scholar : PubMed/NCBI

170 

Stavru F, Hülsmann BB, Spang A, Hartmann E, Cordes VC and Görlich D: NDC1: A crucial membrane-integral nucleoporin of metazoan nuclear pore complexes. J Cell Biol. 173:509–519. 2006. View Article : Google Scholar : PubMed/NCBI

171 

D'Angelo MA and Hetzer MW: Structure, dynamics and function of nuclear pore complexes. Trends Cell Biol. 18:456–466. 2008. View Article : Google Scholar : PubMed/NCBI

172 

Mahipal A and Malafa M: Importins and exportins as therapeutic targets in cancer. Pharmacol Ther. 164:135–143. 2016. View Article : Google Scholar : PubMed/NCBI

173 

Qiao W, Han Y, Jin W, Tian M, Chen P, Min J, Hu H, Xu B, Zhu W, Xiong L and Lin Q: Overexpression and biological function of TMEM48 in non-small cell lung carcinoma. Tumour Biol. 37:2575–2586. 2016. View Article : Google Scholar

174 

Jiang XY, Wang L, Liu ZY, Song WX, Zhou M and Xi L: TMEM48 promotes cell proliferation and invasion in cervical cancer via activation of the Wnt/β-catenin pathway. J Recept Signal Transduct Res. 41:371–377. 2021. View Article : Google Scholar

175 

Nazli A, Safdar A, Saleem A, Akhtar M, Brady LI, Schwartzentruber J and Tarnopolsky MA: A mutation in the TMEM65 gene results in mitochondrial myopathy with severe neurological manifestations. Eur J Hum Genet. 25:744–751. 2017. View Article : Google Scholar : PubMed/NCBI

176 

Zhang B, Liu Q, Wen W, Gao H, Wei W, Tang A, Qin B, Lyu H, Meng X, Li K, et al: The chromatin remodeler CHD6 promotes colorectal cancer development by regulating TMEM65-mediated mitochondrial dynamics via EGF and Wnt signaling. Cell Discov. 8:1302022. View Article : Google Scholar : PubMed/NCBI

177 

Zhang YL, Huang MY, Yang SY, Cai JY, Zhao Q, Zhang FL, Hu X, Shao ZM, Liao L, Cao AY and Li DQ: MYC/TET3-regulated TMEM65 activates OXPHOS-SERPINB3 pathway to promote progression and cisplatin resistance in triple-negative breast cancer. Adv Sci. 12:e004212025. View Article : Google Scholar

178 

Qian Z, Liang J, Huang R, Song W, Ying J, Bi X, Zhao J, Shi Z, Liu W, Liu J, et al: HBV integrations reshaping genomic structures promote hepatocellular carcinoma. Gut. 73:1169–1182. 2024. View Article : Google Scholar : PubMed/NCBI

179 

Song X, Wang P, Feng R, Chetry M, Li E, Wu X, Liu Z, Liao S and Lin J: Pan-cancer analysis of prognostic and immune infiltrates for the TMEM65, especially for the breast cancer. Evid Based Complement Alternat Med. 2023:93494942023. View Article : Google Scholar : PubMed/NCBI

180 

Li X, Zhou J, Zhang W, You W, Wang J, Zhou L, Liu L, Chen WW and Li H: Pan-cancer analysis identifies tumor cell surface targets for CAR-T cell therapies and antibody drug conjugates. Cancers. 14:56742022. View Article : Google Scholar : PubMed/NCBI

181 

Li B, Huang MZ, Wang XQ, Tao BB, Zhong J, Wang XH, Zhang WC and Li ST: TMEM140 is associated with the prognosis of glioma by promoting cell viability and invasion. J Hematol Oncol. 8:892015. View Article : Google Scholar : PubMed/NCBI

182 

Shimizu N, Noda S, Katayama K, Ichikawa H, Kodama H and Miyoshi H: Identification of genes potentially involved in supporting hematopoietic stem cell activity of stromal cell line MC3T3-G2/PA6. Int J Hematol. 87:239–245. 2008. View Article : Google Scholar : PubMed/NCBI

183 

Guan Y, Guo L, Yang E, Liao Y, Liu L, Che Y, Zhang Y, Wang L, Wang J and Li Q: HSV-1 nucleocapsid egress mediated by UL31 in association with UL34 is impeded by cellular transmembrane protein 140. Virology. 464-465:1–10. 2014. View Article : Google Scholar : PubMed/NCBI

184 

Chen WC, Wang CY, Hung YH, Weng TY, Yen MC and Lai MD: Systematic analysis of gene expression alterations and clinical outcomes for long-chain acyl-coenzyme a synthetase family in cancer. PLoS One. 11:e01556602016. View Article : Google Scholar : PubMed/NCBI

185 

Refaat A, Owis M, Abdelhamed S, Saiki I and Sakurai H: Retrospective screening of microarray data to identify candidate IFN-inducible genes in a HTLV-1 transformed model. Oncol Lett. 15:4753–4758. 2018.PubMed/NCBI

186 

Baker Frost D, da Silveira W, Hazard ES, Atanelishvili I, Wilson RC, Flume J, Day KL, Oates JC, Bogatkevich GS, Feghali-Bostwick C, et al: Differential DNA methylation landscape in skin fibroblasts from African Americans with systemic sclerosis. Genes. 12:1292021. View Article : Google Scholar : PubMed/NCBI

187 

Barradas M, Gonos ES, Zebedee Z, Kolettas E, Petropoulou C, Delgado MD, León J, Hara E and Serrano M: Identification of a candidate tumor-suppressor gene specifically activated during Ras-induced senescence. Exp Cell Res. 273:127–137. 2002. View Article : Google Scholar : PubMed/NCBI

188 

Li J, Wang X, Chen L, Zhang J, Zhang Y, Ren X, Sun J, Fan X, Fan J, Li T, et al: TMEM158 promotes the proliferation and migration of glioma cells via STAT3 signaling in glioblastomas. Cancer Gene Ther. 29:1117–1129. 2022. View Article : Google Scholar : PubMed/NCBI

189 

Zirn B, Samans B, Wittmann S, Pietsch T, Leuschner I, Graf N and Gessler M: Target genes of the WNT/beta-catenin pathway in Wilms tumors. Genes Chromosomes Cancer. 45:565–574. 2006. View Article : Google Scholar : PubMed/NCBI

190 

Mohammed Ael S, Eguchi H, Wada S, Koyama N, Shimizu M, Otani K, Ohtaki M, Tanimoto K, Hiyama K, Gaber MS and Nishiyama M: TMEM158 and FBLP1 as novel marker genes of cisplatin sensitivity in non-small cell lung cancer cells. Exp Lung Res. 38:463–474. 2012. View Article : Google Scholar : PubMed/NCBI

191 

Fu Y, Yao N, Ding D, Zhang X, Liu H, Ma L, Shi W, Zhu C and Tang L: TMEM158 promotes pancreatic cancer aggressiveness by activation of TGFβ1 and PI3K/AKT signaling pathway. J Cell Physiol. 235:2761–2775. 2020. View Article : Google Scholar

192 

Cui X, Lu J, Zhao C and Duan Y: Oncogenic transmembrane protein 158 drives the PI3K/Akt signaling pathway to accelerate gastric cancer cell growth. Braz J Med Biol Res. 56:e129432023. View Article : Google Scholar : PubMed/NCBI

193 

Liu L, Zhang J, Li S, Yin L and Tai J: Silencing of TMEM158 inhibits tumorigenesis and multidrug resistance in colorectal cancer. Nutr Cancer. 72:662–671. 2020. View Article : Google Scholar

194 

Cheng Z, Guo J, Chen L, Luo N, Yang W and Qu X: Overexpression of TMEM158 contributes to ovarian carcinogenesis. J Exp Clin Cancer Res. 34:752015. View Article : Google Scholar : PubMed/NCBI

195 

Zhu X, Liu T and Yin X: TMEM158, as plasma cfRNA marker, promotes proliferation and doxorubicin resistance in ovarian cancer. Pharmacogenomics J. 24:342024. View Article : Google Scholar : PubMed/NCBI

196 

Tong J, Li H, Hu Y, Zhao Z and Li M: TMEM158 regulates the canonical and non-canonical pathways of TGF-β to mediate EMT in triple-negative breast cancer. J Cancer. 13:2694–2704. 2022. View Article : Google Scholar :

197 

Li J, Hou H, Sun J, Ding Z, Xu Y and Li G: Systematic pan-cancer analysis identifies transmembrane protein 158 as a potential therapeutic, prognostic and immunological biomarker. Funct Integr Genomics. 23:1052023. View Article : Google Scholar : PubMed/NCBI

198 

Huang J, Liu W, Zhang D, Lin B and Li B: TMEM158 expression is negatively regulated by AR signaling and associated with favorite survival outcomes in prostate cancers. Front Oncol. 12:10234552022. View Article : Google Scholar : PubMed/NCBI

199 

Jawinski P, Kirsten H, Sander C, Spada J, Ulke C, Huang J, Burkhardt R, Scholz M, Hensch T and Hegerl U: Human brain arousal in the resting state: A genome-wide association study. Mol Psychiatry. 24:1599–1609. 2019. View Article : Google Scholar

200 

Goodman JM: LDAF1 holds the key to seipin function. Dev Cell. 51:544–545. 2019. View Article : Google Scholar : PubMed/NCBI

201 

Shi J, Zhang Y, Chen Y, Guo T and Piao H: Identification of TMEM159 as a biomarker of glioblastoma progression based on immune characteristics. Biocell. 48:1241–1263. 2024. View Article : Google Scholar

202 

Geng F, Cheng X, Wu X, Yoo JY, Cheng C, Guo JY, Mo X, Ru P, Hurwitz B, Kim SH, et al: Inhibition of SOAT1 suppresses glioblastoma growth via blocking SREBP-1-mediated lipogenesis. Clin Cancer Res. 22:5337–5348. 2016. View Article : Google Scholar : PubMed/NCBI

203 

Castro IG, Eisenberg-Bord M, Persiani E, Rochford JJ, Schuldiner M and Bohnert M: Promethin is a conserved seipin partner protein. Cells. 8:2682019. View Article : Google Scholar : PubMed/NCBI

204 

Bohnert M: New friends for seipin - Implications of seipin partner proteins in the life cycle of lipid droplets. Semin Cell Dev Biol. 108:24–32. 2020. View Article : Google Scholar : PubMed/NCBI

205 

Chartschenko E, Hugenroth M, Akhtar I, Droste A, Kolkhof P, Bohnert M and Beller M: CG32803 is the fly homolog of LDAF1 and influences lipid storage in vivo. Insect Biochem Mol Biol. 133:1035122021. View Article : Google Scholar

206 

Wagner V, Morton M, Dorayappan KDP, Gonzalez A, Yu L, Sakaue T, Conrads T, Maxwell GL, Cosgrove C, Backes F, et al: Circulating extracellular vesicles protein expression for early prediction of platinum-resistance in high-grade serous ovarian cancer. Oncogene. 44:1197–1203. 2025. View Article : Google Scholar : PubMed/NCBI

207 

Rao J, Wu X, Zhou X, Deng R and Ma Y: TMEM205 is an independent prognostic factor and is associated with immune cell infiltrates in hepatocellular carcinoma. Front Genet. 11:5757762020. View Article : Google Scholar : PubMed/NCBI

208 

Saini U, Smith BQ, Dorayappan KDP, Yoo JY, Maxwell GL, Kaur B, Konishi I, O'Malley D, Cohn DE and Selvendiran K: Targeting TMEM205 mediated drug resistance in ovarian clear cell carcinoma using oncolytic virus. J Ovarian Res. 15:1302022. View Article : Google Scholar : PubMed/NCBI

209 

Calo CA, Smith BQ, Dorayappan KDP, Saini U, Lightfoot M, Wagner V, Kalaiyarasan D, Cosgrove C, Wang QE, Maxwell GL, et al: Aberrant expression of TMEM205 signaling promotes platinum resistance in ovarian cancer: An implication for the antitumor potential of DAP compound. Gynecol Oncol. 164:136–145. 2022. View Article : Google Scholar

210 

Sen T, Rodriguez BL, Chen L, Corte CMD, Morikawa N, Fujimoto J, Cristea S, Nguyen T, Diao L, Li L, et al: Targeting DNA damage response promotes antitumor immunity through STING-Mediated T-cell activation in small cell lung cancer. Cancer Discov. 9:646–661. 2019. View Article : Google Scholar : PubMed/NCBI

211 

Jiang M, Chen P, Wang L, Li W, Chen B, Liu Y, Wang H, Zhao S, Ye L, He Y and Zhou C: cGAS-STING, an important pathway in cancer immunotherapy. J Hematol Oncol. 13:812020. View Article : Google Scholar : PubMed/NCBI

212 

Pantelidou C, Sonzogni O, De Oliveria Taveira M, Mehta AK, Kothari A, Wang D, Visal T, Li MK, Pinto J, Castrillon JA, et al: PARP inhibitor efficacy depends on CD8(+) T-cell recruitment via intratumoral STING pathway activation in BRCA-Deficient models of triple-negative breast cancer. Cancer Discov. 9:722–737. 2019. View Article : Google Scholar : PubMed/NCBI

213 

Ni H, Zhang H, Li L, Huang H, Guo H, Zhang L, Li C, Xu JX, Nie CP, Li K, et al: T cell-intrinsic STING signaling promotes regulatory T cell induction and immunosuppression by upregulating FOXP3 transcription in cervical cancer. J Immunother Cancer. 10:e0051512022. View Article : Google Scholar : PubMed/NCBI

214 

Zhang P, Rashidi A, Zhao J, Silvers C, Wang H, Castro B, Ellingwood A, Han Y, Lopez-Rosas A, Zannikou M, et al: STING agonist-loaded, CD47/PD-L1-targeting nanoparticles potentiate antitumor immunity and radiotherapy for glioblastoma. Nat Commun. 14:16102023. View Article : Google Scholar : PubMed/NCBI

215 

Lam KC, Araya RE, Huang A, Chen Q, Di Modica M, Rodrigues RR, Lopès A, Johnson SB, Schwarz B, Bohrnsen E, et al: Microbiota triggers STING-type I IFN-dependent monocyte reprogramming of the tumor microenvironment. Cell. 184:5338–5356.e21. 2021. View Article : Google Scholar : PubMed/NCBI

216 

Wu J, Chen YJ, Dobbs N, Sakai T, Liou J, Miner JJ and Yan N: STING-mediated disruption of calcium homeostasis chronically activates ER stress and primes T cell death. J Exp Med. 216:867–883. 2019. View Article : Google Scholar : PubMed/NCBI

217 

Zhang G, Shu Z, Yu J, Li J, Yi P, Wu B, Deng D, Yan S, Li Y, Ren D, et al: High ANO1 expression is a prognostic factor and correlated with an immunosuppressive tumor microenvironment in pancreatic cancer. Front Immunol. 15:13412092024. View Article : Google Scholar : PubMed/NCBI

218 

Jiang F, Jia K, Chen Y, Ji C, Chong X, Li Z, Zhao F, Bai Y, Ge S, Gao J, et al: ANO1-mediated inhibition of cancer ferroptosis confers immunotherapeutic resistance through recruiting cancer-associated fibroblasts. Adv Sci. 10:e23008812023. View Article : Google Scholar

219 

Kuang L, Pang Y and Fang Q: TMEM101 expression and its impact on immune cell infiltration and prognosis in hepatocellular carcinoma. Sci Rep. 14:318472024. View Article : Google Scholar

220 

Lee Y, Ko D, Yoon J and Kim S: TMEM52B-derived peptides inhibit generation of soluble E-cadherin and EGFR activity to suppress colon cancer growth and early metastasis. J Transl Med. 23:1462025. View Article : Google Scholar : PubMed/NCBI

221 

Sermeus A, Cosse JP, Crespin M, Mainfroid V, de Longueville F, Ninane N, Raes M, Remacle J and Michiels C: Hypoxia induces protection against etoposide-induced apoptosis: Molecular profiling of changes in gene expression and transcription factor activity. Mol Cancer. 7:272008. View Article : Google Scholar : PubMed/NCBI

222 

Berezin IV, Kershengol'ts BM and Ugarova NN: Microenvironment of enzymes as 1 of the factors determining enzyme stability. Stabilization of soluble and immobilized horseradish peroxidase. Dokl Akad Nauk SSSR. 223:1256–1259. 1975.In Russian. PubMed/NCBI

223 

Lee PS, Teaberry VS, Bland AE, Huang Z, Whitaker RS, Baba T, Fujii S, Secord AA, Berchuck A and Murphy SK: Elevated MAL expression is accompanied by promoter hypomethylation and platinum resistance in epithelial ovarian cancer. Int J Cancer. 126:1378–1389. 2010. View Article : Google Scholar

224 

Lee H and Evans T: TMEM88 Inhibits Wnt signaling by promoting Wnt signalosome localization to multivesicular bodies. iScience. 19:267–280. 2019. View Article : Google Scholar : PubMed/NCBI

225 

Shen DW and Gottesman MM: RAB8 enhances TMEM205-mediated cisplatin resistance. Pharm Res. 29:643–650. 2012. View Article : Google Scholar :

226 

Shen DW, Ma J, Okabe M, Zhang G, Xia D and Gottesman MM: Elevated expression of TMEM205, a hypothetical membrane protein, is associated with cisplatin resistance. J Cell Physiol. 225:822–828. 2010. View Article : Google Scholar : PubMed/NCBI

227 

Fu Q, Wu X, Lu Z, Chang Y, Jin Q, Jin T and Zhang M: TMEM205 induces TAM/M2 polarization to promote cisplatin resistance in gastric cancer. Gastric Cancer. 27:998–1015. 2024. View Article : Google Scholar : PubMed/NCBI

228 

Hu Y, Zhang Y, He J, Rao H, Zhang D, Shen Z and Zhou C: ANO1: central role and clinical significance in non-neoplastic and neoplastic diseases. Front Immunol. 16:15703332025. View Article : Google Scholar : PubMed/NCBI

229 

Vyas A, Duvvuri U and Kiselyov K: Copper-dependent ATP7B up-regulation drives the resistance of TMEM16A-overexpressing head-and-neck cancer models to platinum toxicity. Biochem J. 476:3705–3719. 2019. View Article : Google Scholar : PubMed/NCBI

230 

Carvalho V, Pronk JW and Engel AH: Characterization of membrane proteins using cryo-electron microscopy. Curr Protoc Protein Sci. 94:e722018. View Article : Google Scholar : PubMed/NCBI

231 

Duan J, He XH, Li SJ and Xu HE: Cryo-electron microscopy for GPCR research and drug discovery in endocrinology and metabolism. Nat Rev Endocrinol. 20:349–365. 2024. View Article : Google Scholar : PubMed/NCBI

232 

Robertson MJ, Meyerowitz JG and Skiniotis G: Drug discovery in the era of cryo-electron microscopy. Trends Biochem Sci. 47:124–135. 2022. View Article : Google Scholar

233 

Chang A, Xiang X, Wang J, Lee C, Arakhamia T, Simjanoska M, Wang C, Carlomagno Y, Zhang G, Dhingra S, et al: Homotypic fibrillization of TMEM106B across diverse neurodegenerative diseases. Cell. 185:1346–1355.e15. 2022. View Article : Google Scholar : PubMed/NCBI

234 

Yuan Y, Kong F, Xu H, Zhu A, Yan N and Yan C: Cryo-EM structure of human glucose transporter GLUT4. Nat Commun. 13:26712022. View Article : Google Scholar : PubMed/NCBI

235 

Shang G, Zhang C, Chen ZJ, Bai XC and Zhang X: Cryo-EM structures of STING reveal its mechanism of activation by cyclic GMP-AMP. Nature. 567:389–393. 2019. View Article : Google Scholar : PubMed/NCBI

236 

Zhang X, Wu C, Wei T, Lu Y, Liu C and Zhang J: Cryo-EM studies of the apo states of human IGF1R. Biochem Biophys Res Commun. 618:148–152. 2022. View Article : Google Scholar : PubMed/NCBI

237 

García-Nafría J and Tate CG: Cryo-electron microscopy: Moving beyond X-ray crystal structures for drug receptors and drug development. Annu Rev Pharmacol Toxicol. 60:51–71. 2020. View Article : Google Scholar

238 

Khorn PA, Luginina AP, Pospelov VA, Dashevsky DE, Khnykin AN, Moiseeva OV, Safronova NA, Belousov AS, Mishin AV and Borshchevsky VI: Rational design of drugs targeting G-Protein-coupled receptors: A structural biology perspective. Biochemistry. 89:747–764. 2024.PubMed/NCBI

239 

Madej MG and Ziegler CM: Dawning of a new era in TRP channel structural biology by cryo-electron microscopy. Pflugers Arch. 470:213–225. 2018. View Article : Google Scholar : PubMed/NCBI

240 

Sung MW, Hu K, Hurlimann LM, Lees JA, Fennell KF, West MA, Costales C, Rodrigues AD, Zimmermann I, Dawson RJP, et al: Cyclosporine A sterically inhibits statin transport by solute carrier OATP1B1. J Biol Chem. 301:1084842025. View Article : Google Scholar : PubMed/NCBI

241 

Ansell TB, Song W, Coupland CE, Carrique L, Corey RA, Duncan AL, Cassidy CK, Geurts MMG, Rasmussen T, Ward AB, et al: LipIDens: Simulation assisted interpretation of lipid densities in cryo-EM structures of membrane proteins. Nat Commun. 14:77742023. View Article : Google Scholar : PubMed/NCBI

242 

Ackle F, Thavarasah S, Earp JC and Seeger MA: Rigid enlargement of sybodies with antibody fragments for cryo-EM analyses of small membrane proteins. Sci Rep. 15:94602025. View Article : Google Scholar : PubMed/NCBI

243 

Sun C and Gennis RB: Single-particle cryo-EM studies of transmembrane proteins in SMA copolymer nanodiscs. Chem Phys Lipids. 221:114–119. 2019. View Article : Google Scholar : PubMed/NCBI

Related Articles

  • Abstract
  • View
  • Download
  • Twitter
Copy and paste a formatted citation
Spandidos Publications style
Shi J, Zheng D, Yao B, Liu Q, Xu H and Piao H: Research progress on TMEM proteins in cancer progression and chemoresistance (Review). Int J Mol Med 56: 219, 2025.
APA
Shi, J., Zheng, D., Yao, B., Liu, Q., Xu, H., & Piao, H. (2025). Research progress on TMEM proteins in cancer progression and chemoresistance (Review). International Journal of Molecular Medicine, 56, 219. https://doi.org/10.3892/ijmm.2025.5660
MLA
Shi, J., Zheng, D., Yao, B., Liu, Q., Xu, H., Piao, H."Research progress on TMEM proteins in cancer progression and chemoresistance (Review)". International Journal of Molecular Medicine 56.6 (2025): 219.
Chicago
Shi, J., Zheng, D., Yao, B., Liu, Q., Xu, H., Piao, H."Research progress on TMEM proteins in cancer progression and chemoresistance (Review)". International Journal of Molecular Medicine 56, no. 6 (2025): 219. https://doi.org/10.3892/ijmm.2025.5660
Copy and paste a formatted citation
x
Spandidos Publications style
Shi J, Zheng D, Yao B, Liu Q, Xu H and Piao H: Research progress on TMEM proteins in cancer progression and chemoresistance (Review). Int J Mol Med 56: 219, 2025.
APA
Shi, J., Zheng, D., Yao, B., Liu, Q., Xu, H., & Piao, H. (2025). Research progress on TMEM proteins in cancer progression and chemoresistance (Review). International Journal of Molecular Medicine, 56, 219. https://doi.org/10.3892/ijmm.2025.5660
MLA
Shi, J., Zheng, D., Yao, B., Liu, Q., Xu, H., Piao, H."Research progress on TMEM proteins in cancer progression and chemoresistance (Review)". International Journal of Molecular Medicine 56.6 (2025): 219.
Chicago
Shi, J., Zheng, D., Yao, B., Liu, Q., Xu, H., Piao, H."Research progress on TMEM proteins in cancer progression and chemoresistance (Review)". International Journal of Molecular Medicine 56, no. 6 (2025): 219. https://doi.org/10.3892/ijmm.2025.5660
Follow us
  • Twitter
  • LinkedIn
  • Facebook
About
  • Spandidos Publications
  • Careers
  • Cookie Policy
  • Privacy Policy
How can we help?
  • Help
  • Live Chat
  • Contact
  • Email to our Support Team