Skip to main content
Erschienen in: Journal of Translational Medicine 1/2023

Open Access 01.12.2023 | Review

Phase separation in cancer at a glance

verfasst von: Qingqing Xie, Jiejuan Cheng, Wuxuan Mei, Dexing Yang, Pengfei Zhang, Changchun Zeng

Erschienen in: Journal of Translational Medicine | Ausgabe 1/2023

Abstract

Eukaryotic cells are segmented into multiple compartments or organelles within the cell that regulate distinct chemical and biological processes. Membrane-less organelles are membrane-less microscopic cellular compartments that contain protein and RNA molecules that perform a wide range of functions. Liquid–liquid phase separation (LLPS) can reveal how membrane-less organelles develop via dynamic biomolecule assembly. LLPS either segregates undesirable molecules from cells or aggregates desired ones in cells. Aberrant LLPS results in the production of abnormal biomolecular condensates (BMCs), which can cause cancer. Here, we explore the intricate mechanisms behind the formation of BMCs and its biophysical properties. Additionally, we discuss recent discoveries related to biological LLPS in tumorigenesis, including aberrant signaling and transduction, stress granule formation, evading growth arrest, and genomic instability. We also discuss the therapeutic implications of LLPS in cancer. Understanding the concept and mechanism of LLPS and its role in tumorigenesis is crucial for antitumor therapeutic strategies.
Hinweise

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Abkürzungen
LLPS
Liquid–liquid phase separation
TCR
T-cell receptor
ZAP70
Zeta chain of TCR-associated protein kinase 70
GRB2
Growth factor receptor-bound protein 2
PLC
Phospholipase C
GADs
GRB2-related adaptor downstream of Shc
SLP76
SH2 domain-containing leukocyte protein of 76 kDa
NCK
Noncatalytic region of tyrosine kinase
N‑WASP
Neural Wiskott-Aldrich syndrome protein
ARP2/3
Actin-related protein 2/3
RTKs
Receptor tyrosine kinases
EML4
Echinoderm microtubule-associated protein-like 4
ALK
Anaplastic lymphoma kinase
CCDC6
Coiled-coil domain-containing protein 6
RET
Rearranged during transfection
SOS1
Son of sevenless 1
SHP2
SH2 domain-containing protein tyrosine phosphatase 2
cAMP
3′,5′-Cyclic adenosine monophosphate
PKA
Protein kinase A
DnaJB1
DnaJ Hsp40 member B1
G6PC
Glucose 6-phosphatase
MST1/2
Laforin-macrophage stimulating 1/2
DACT1
Disheveled binding antagonist of -catenin 1
TGF-
Transforming growth factor
CK2
Sequester casein kinase 2
SGs
Stress granules
DEAD
Conserved motif Asp–Glu–Ala–Asp
SPOP
Speckle-type pox virus and zinc finger (POZ) protein
CRL3
Cullin-RING E3 ligase
NLS
Nuclear localization sequence
DAXX
Death domain-associated protein
MYC
Myelocytomatosis oncogene
AR
Androgen receptor
GLI3
GLI family zinc finger 3
SRC3
Steroid receptor coactivator 3
PML
Promyelocytic leukemia
DBD
DNA-binding domain
CTD
C-terminal regulatory domain
OD
Oligomerization domain
UBR
Unstructured basic region
TD
Tetramerization domain
DDR
DNA damage response
PAR
Poly (ADP-ribose)
PARP1
Poly (ADP-ribose) polymerase 1
LCD
Low complexity domain
DSB
DNA double-strand break
CycT1
Cyclin T1
PROTAC
Proteolysis-targeting chimera
PTMs
Posttranslational modifications
RNP
Ribonucleoprotein
UBQLN2
UBA domain of ubiquitin 2
PAS
Pre-autophagosomal structure
ER
Estrogen receptor
Tau-5
Transduction unit 5 region

Introduction

The execution of biological functions depends on the coordination of complex biochemical events in a dense cellular space. Cancer is known to cause widespread disruption to the regulation of various cellular processes. These processes include transcription, chromatin organization, RNA processing, genomic integrity, and signaling, which are responsible for diseases characterized by unregulated cell proliferation and growth, enhanced cell survival, and metabolic reprogramming [1]. The molecular mechanism by which coordination occurs in biological pathways may shed light on the molecular basis of cancerous pathway disruption. Eukaryotic cells evolved unique inner structures, compartments, or organelles with specific properties and functions to orchestrate biochemical reactions. Typically, lipid bilayer membranes separate the exterior from the interior environment of a membrane-bound organelle, such as the nucleus, endoplasmic reticulum, lysosomes, mitochondria, secretory vesicles, and Golgi apparatus [2].
In eukaryotic cells, biomolecular condensates (BMCs) are membrane-less assemblies that play a crucial role in many cellular functions via the compartmentalization of specific nucleic acids and proteins within subcellular compartments. Membrane-less bodies, including leukopoietins, nucleosomes, parameres, and Cajal bodies, stress granules (SGs), signaling sites, and processing bodies, are involved in signaling, punctum formation, and transduction, ribosome biogenesis, and cell division [3]. The formation of BMCs can be attributed to weak, multivalent interactions between macromolecules such as proteins and nucleic acids [4, 5]. Increasing evidence suggests that BMCs are dynamically and reversibly assembled through liquid–liquid phase separation (LLPS) [6]. Various components of functional membrane-less organelles exist as liquid-like droplets [7]. Cells have evolved multiple mechanisms to maintain LLPS. Intrinsically disordered regions (IDRs), proteins with multiple folded domains, metal ion-controlled RNA-binding domains and nucleic acid chains can drive LLPS through their multivalency [6, 8]. Metal ions are also involved in the formation of BMCs through LLPS [8]. A dysfunctional LLPS leads to tumorigenesis through the disruption of BMCs [9]. Here, we summarize recent insights and findings on the formation, regulation, function, and mechanism of LLPS. In addition, we present how LLPS contributes to the development of cancer. Finally, we describe how innovative cancer therapies may benefit from the regulation of the LLPS.

Biological characteristics of LLPS

LLPS occurs within cells as macromolecules transition between dense and diluted phases, which is necessary for achieving the lowest free energy state [10]. A characteristic of LLPS is the presence of threshold concentrations of macromolecules, which are affected by several biophysical parameters, including temperature, salt concentration, and ions [6]. Proteins, DNA, and RNA, which are the primary elements and mediators of LLPS, may interact to create a highly multi-component system. Multivalent interactions exhibit high affinity and stereospecificity, which enables the assembly of oligomers and polymers. A higher valency allows for the development of larger oligomers or polymers at lower saturation levels. Nucleic acid chains, IDRs, and multiple folded domains can all result in multivalency in proteins [11]. The LLPS process can produce several states, including liquid droplets, hydrogels, and fibrous aggregates. In some cases, the aggregation state can affect the protein function [3]. Scaffold molecules are composed of multivalent nucleic acids or proteins that are critical for the initiation of LLPS. The surfaces of macromolecules will be subjected to transient, weak, nonspecific interactions with their counterparts in the cytoplasm or nucleoplasm [12]. Many structural characteristics are observed in proteins undergoing LLPS, such as repetitive modular domains, weakly adhesive multivalent motifs, IDRs, nucleic acid recognition domains, and oligomerization or dimerization domains [13]. Many RNA-binding proteins possess IDRs and low complexity domains (LCDs), known as prion-like domains, which allow the formation of BMCs in a highly crowded nuclear environment [14]. BMCs can be facilitated by weak interactions, such as hydrophobic, electrostatic, cation–pi, and pi–pi interactions [15].
Protein-dependent LLPSs are biophysical processes that rely heavily on scaffolds and clients for their formation and function, providing a mechanism for regulating and compartmentalizing biochemical processes inside cells (Fig. 1) [9]. Only a few LCDs can maintain a three-dimensional structure, and these regions serve as scaffolds in motif interaction. Typically, the disordered region is highly enriched in specific amino acids containing aromatic residues, charged residues, or hydrophilic residues, all of which are associated with pi–pi, cation–Pi, and electrostatic interactions among these amino acids. Conversely, aliphatic residues such as isoleucine, leucine, and valine are less common in LCDs [16]. Scaffold molecules in cells are typically highly concentrated and have multiple valences based on the number of disturbing modules [17, 18]. Additionally, post-translational modifications of scaffolds can control the condensate assembly by modulating weak interactions between multivalent molecules that lead to LLPS. The type and state of the scaffold protein, the recruited client cells, and their location can impact the physiological functions of BMCs [17]. A variety of forces are simultaneously involved in motif interactions and driving LLPS, such as sheets, electrostatic interactions (cation-anion), Van der Waals forces (dipole–dipole), and -effects (cation–, ‑stacking) [13, 17, 19]. Proteins with repetitive sequences of proline-rich motifs and Src homology 3 domains can also serve as endogenous LLPS through concentration-dependent mechanisms. RNA can participate in the cellular development of RNA/protein-rich membrane-less aggregates by facilitating LLPS. A low ratio of RNA to protein can facilitate LLPS droplet formation, whereas a high RNA to protein ratio can inhibit it. A reduction in nuclear RNA levels or a genetic alteration that affects RNA binding within cells can result in LLPS, which may cause aggregate formation [19].

Liquid–liquid phase separation in cancer

In recent decades, tremendous progress has been made in understanding malignant tumors. Cancer cells are characterized by several traits, including sustained proliferative signaling, resistance to cell death, induction of angiogenesis, replication of immortality, evasion of growth suppressors, and activation of invasion and metastasis [20]. In addition, new hallmarks of cancer have emerged, including the deregulation of cellular metabolism, genome instability, mutation, senescent cells, polymorphic microbiomes, tumor-promoting inflammation, replication immortality, immunity to destruction, and non-mutational epigenetic reprogramming [21]. Despite this, the mechanism by which cancer develops remains unclear. Numerous studies have established the crucial contribution of LLPS to cancer, which regulates various biological processes such as X-chromosome inactivation/paraspeckle formation, transcription/chromatin organization, cytoplasmic DNA sensing, DNA damage response, SGs formation, proteasome/autophagosome formation, tumorigenesis, synaptic vesicle active zone formation, and ribonucleoprotein (RNP) synthesis (Fig. 2) [22]. LLPS provides new perspectives for understanding cancer development and may facilitate the development of potential therapeutic strategies.

Misregulated signaling and transduction

Cancer-related mutations can trigger the formation of BMCs. LLPS facilitates signaling condensate formation and modulates immune signaling pathways that include the downstream of B cell receptor, T cell receptor, as well as innate immune receptors such as retinoic acid-inducible gene I protein (RIG-I) and cyclic GMP-AMP synthase (cGAS)-stimulator of interferon genes (STING) [23]. Neurofibromin 2 (NF2) stimulates innate immunity through the regulation of YAP/TAZ-mediated inhibition of tank-binding kinase 1 (TBK1). NF2 mutants suppress cyclic GMP-AMP synthase (cGAS)-stimulator of interferon genes (STING) and RIG-I-like Receptor (RLR)-mitochondrial antiviral signaling protein (MAVS) signaling. LLPS of NF2 mutants impairs antitumor immunity induced by STING [24]. 2′3′-cGAMP (cyclic GMP-AMP) causes the LLPS of endoplasmic reticulum (ER)-associated STING to generate BMCs. STING BMCs restrict the overactivation of innate immunity through the separation of STING-TBK1 and IRF3 [25]. As a result of LLPS, membrane receptors and their signaling molecules can self-assemble during signaling, such as during T-cell receptor (TCR) signaling in immune cells. In response to TCR activation, tyrosine residues in LA were phosphorylated by the zeta chain of TCR-associated protein kinase 70 (ZAP70) [26]. As a result, the formation of condensation is induced by the attraction of multivalent SH-containing proteins, such as growth factor receptor-bound protein 2 (GRB2), phospholipase C (PLC), and GRB2-related adaptor downstream of Shc (GADs). Following this, SOS1 is recruited to participate in the activation of RAS signaling [27]. Furthermore, the SH2 domain-containing leukocyte protein of 76 kDa (SLP76) attaches to GRB2 or GADs and recruits the noncatalytic region of tyrosine kinase (NCK) and neural Wiskott-Aldrich syndrome protein (N‑WASP) actin effectors and the actin-related protein 2/3 (ARP2/3) complex that is necessary to assemble actin filaments. LLPS prevents spontaneous SOS1 membrane localization from membrane-dependent RAS activation, thus initiating RAS signaling. Nevertheless, oncoproteins of the chimeric receptor tyrosine kinases (RTKs) impair RAS signaling mediated by LLPS [26, 28].
Mutant oncoproteins exhibit LLPS due to their multivalent interactions. In the case of echinoderm microtubule-associated protein-like 4 (EML4)-anaplastic lymphoma kinase (ALK) or coiled-coil domain-containing protein 6 (CCDC6)-rearranged during transfection (RET), multimerization domains of EML4 or CCDC6 were assembled, whereas targeting sequences of ALK or RET were lost, resulting in the production of granules without membranes [13, 29]. As a result, these granules can create concentrations of the RAS-activating complex GRB2/son of sevenless 1 (SOS1), thereby facilitating RAS signaling independent of the membrane lipids. LLPS is enabled by mutant SH2 domain-containing protein tyrosine phosphatase 2 (SHP2) with an open conformation in its PTP domain. Oncogenic SHP2 mutants activate RAS-MAP kinase signaling through LLPS by recruiting wild-type SHP2 into BMCs [30].
The 3′,5′-cyclic adenosine monophosphate (cAMP)-dependent protein kinase A (PKA) signaling transduction further reveals that LLPS facilitates the rapid concentration of critical signaling pathway components in the cytoplasm in cancer cells. PKA, a tetrameric holoenzyme, is composed of a dimer of regulatory subunits and a pair of catalytic subunits. The non-redundant regulatory subunit RI, a cAMP receptor, is essential for the function of PKA [31, 32]. In normal cells, the RI can contribute to BMCs that can sequester high concentrations of cAMP and maintain high levels of PKA activity. It can also function as a dynamic buffer for cAMP. In atypical liver fibrolamellar carcinoma, the DnaJ Hsp40 member B1 (DnaJB1)-PKAcat fusion proteins replace the native N terminus of PKA-C with the J domain of DnaJB1, which abolishes the RI LLPS, enhances the levels of cAMP in positioned phosphodiesterase sinks, and activates cAMP signaling [31].
Abnormal LLPS are also associated with the dysfunction of signal transduction pathways. Glycogen accumulation occurs in precancerous liver lesions due to the downregulation of the glucose 6-phosphatase (G6PC), which modulates hepatic glycogenolysis. The LLPS then forms condensed glycogen compartments from the accumulated glycogen. Laforin-macrophage stimulating 1/2 (MST1/2) complexes aggregate in glycogen-liquid droplets to relieve MST1/2 inhibition on the yes-associated protein (YAP), enabling tumorigenic cell transformation. The disheveled binding antagonist of -catenin 1 (DACT1), a target gene of transforming growth factor (TGF-), can assemble BMCs within the cytoplasm that sequester casein kinase 2 (CK2) and prevent WNT signaling. Thus, DACT1 BMCs play a vital role in the metastatic process of breast and prostate cancer cells [33, 34]. IDRs within the signalosome scaffold protein Dishevelled 2 (Dvl2) mediate LLPS. The receptor Fzd5, which is a component of the signalosome, is responsible for stimulating LLPS. Dvl2 LLPS can recruit Axin to stabilize β-catenin, which is crucial for the assembly of the Wnt signalosome and the disassembly of the destruction complex [35, 36].

Stress granules

Under various stress conditions, eukaryotic cells form SGs that halt translation and release mRNA molecules from polysomes. The SGs are cytoplasmic compartments of the RNA granule family that respond to various stress signals and form cytoplasmic BMCs in eukaryotic cells, thereby facilitating cell survival. The assembly of SGs is related to numerous RNA-binding proteins implicated in RNA processing and nuclear transport. Primary SGs are coordinated by the complexes of RNA-binding proteins and mRNA via LLPS [37]. The SG protein G3BP1 enters an auto-inhibited condition in the absence of stress by interacting with its IDRs and arginine-rich regions. Stress conditions cause unfolded mRNAs to liberate G3BP1 from its auto-inhibited conformation, leading to G3BP1 clustering via the interactions between protein and RNA. Subsequently, G3BP1/RNA-phase separated BMCs occurred, and G3BP1 hindered RNA entanglement, recruiting client proteins to facilitate the assembly of SGs [38]. An interaction network between proteins and RNA is responsible for SGs via LLPS. G3BP1 activates RNA-dependent LLPS in response to an increase in the concentration of free RNA. Phosphorylation of IDRs in G3BP1 regulates the formation of LLPS. YB1, a member of the cold shock domain (CSD) family, modulates G3BP1 translation to stimulate SG assembly and cancer metastasis [39]. Furthermore, several types of cancer are associated with DDX3X, a component of SGs, which is an ATP-dependent RNA helicase with the conserved motif Asp–Glu–Ala–Asp (DEAD) [40]. Mutant KRAS tumor cells exhibited a significant increase in SG formation [41, 42]. Extrinsic proteins that interact with G3BP1 can regulate the SGs network. The assembly of RNP SGs is also mediated by heterotypic multivalent interactions [41]. UBQLN2, a shuttle protein for the proteasome, possesses IDRs and is located within the SGs. The oligomerization of UBQLN2 is essential for the occurrence of LLPS. UBQLN2 interacts with ubiquitinated client proteins to facilitate the shuttling of client proteins out of SGs and reverses the UBQLN2-induced LLPS [43].

Evading growth arrest

Evading growth arrest caused by endogenous tumor suppressors can boost cell growth. There is considerable evidence that the nuclear speckle-type pox virus and zinc finger (POZ) protein (SPOP), which serves as a substrate-recognition component of the cullin-RING E3 ligase (CRL3), appears to play a crucial role in carcinogenesis and cancer development [44]. SPOP is a tumor suppressor protein with an N-terminal MATH domain, an internal BTB domain, and a nuclear localization sequence (NLS) at the C-terminus. Several oncoproteins, such as death domain-associated protein (DAXX), myelocytomatosis oncogene (MYC), androgen receptor (AR), GLI family zinc finger 3 (GLI3), and steroid receptor coactivator 3 (SRC3), can be recruited by SPOP to CRL3 for ubiquitination and degradation by proteasomes [45, 46]. The NLS of SPOP facilitates its localization to nuclear speckles, which are RNA-protein granules implicated in the regulation of gene expression, promyelocytic leukemia (PML) bodies, DNA damage loci, metabolism, and splicing. SPOP can oligomerize to localize nuclear speckles, which increases the efficiency of ubiquitination. SPOP interaction with death domain-associated protein (DAXX) can induce droplet formation through LLPS. SPOP alterations impair LLPS and DAXX ubiquitination, which may lead to oncoprotein accumulation [19, 46, 47]. In membrane-less organelles, mutations in the SPOP lead to ubiquitin-dependent protein homeostasis by disrupting LLPS and colocalization [48].
Studies in vivo indicate that p53 participates in cytosolic droplets, implying that LLPS may be critical for the biological activity of p53 [49]. It has been established that the transcriptional regulator p53 regulates processes, including the cell cycle, apoptosis, and senescence, in response to a wide range of stress signals, such as hypoxia, DNA damage, and oxidative stress [50]. The p53 consists of two transactivation domains (TADs), a sequence-specific DNA-binding domain (DBD), a C-terminal regulatory domain (CTD) associated with nuclear localization, and an oligomerization domain (OD) essential for transcriptional activity. The modulation of droplet formation by truncated TP53 mutant points to the significance of multivalent electrostatic interactions between the N-terminal and C-terminal domains of p53 [49]. P53 BMCs depend on the disordered unstructured basic regions (UBRs) that are regulated by electrostatic and hydrophobic interactions. The tetramerization domain (TD) mutations interfere with LLPS of p53 by blocking the production of tetramers [51]. Oncogenic mutations in TD are responsible for preventing the production of p53 BMCs, which results in decreased target gene activation and accelerates cancer development. The disordered TAD of p53 can regulate LLPS and amyloid aggregation. The DBD of p53 underwent LLPS in the presence of polyethylene glycol. DBD mutants of p53, such as M237I and R249S, also experienced LLPS [52].

Maintenance of genome stability

DNA damage events triggered by exogenous and endogenous factors may affect genomic instability and cause cancer. DNA damage response (DDR) and DNA repair processes are frequently disrupted by genetic alterations. DNA repair lesions generate transient and reversible BMCs that induce the repair of proteins and the generation of repair signals. A nucleic acid-like protein modification known as poly (ADP-ribose) (PAR) can be found in BMCs, such as DNA repair foci and SGs. LCDs in PAR-binding or PAR-conjugated proteins promote LLPS and the formation of BMCs. PAR promotes LLPS of proteins with LCDs such as TDP-43, FUS, and hnRNPA1 [53].
The poly (ADP-ribose) polymerase 1 (PARP1) is a critical component of the cancer biology process through its involvement in replication, transcription, chromatin remodeling, genome maintenance, and DNA repair [54]. As a poly (ADP-ribosylation) (PARylation) “writer”, PARP1 synthesizes negatively charged PAR chain polymers. As a result of DNA damage, PARylation is believed to control the physicochemical characteristics, assembly, and catalysis of target proteins [55]. The cellular response to DNA damage is characterized by an increase in PAR levels induced by PARP enzyme hyperactivation at sites of DNA damage, leading to the rapid accumulation of proteins with LCDs. LCD-containing proteins experience LLPS and liquid demixing, which allows cells to filter molecular interactions involving damaged chromatin. Upon dissolution of PAR-seeded liquid compartments, interactions unfold on lesion-flanking chromatin, allowing 53BP1 to accumulate. As a modulator of DNA double-strand break (DSB) repair, 53BP1 stimulates the development of chromatin domains surrounding the damaged DNA [56]. A disruption of LLPS of 53BP1 reduces the 53BP1-dependent activation of p53 and weakens the expression of p53 target genes. During LLPS of 53BP1, localized recognition of DNA damage and repair factor assembly is coordinated with p53-dependent gene activation and cell fate determination [57]. Additionally, 53BP1 promotes the integrity of heterochromatin and genome stability through LLPS in addition to its role in DSB repair [58]. By inactivating P-TEFb, a heterodimer of CDK9 and cyclin T1 (CycT1), PARP1 suppresses Pol II elongation. In response to damage, stimulated PARP1 attaches to transcriptionally active P-TEFb and modifies CycT1 at multiple sites of the protein. P-TEFb-PARP1 signaling contributes to the maintenance of genomic stability and the regulation of transcription quality following DNA damage [59].

Tumor viruses

Evidence suggests that tumor-associated viruses, such as Epstein–Barr virus (EBV) and Kaposi sarcoma herpesvirus (KSHV), regulate the progression of tumors through LLPS [6062]. KSHV genomes interact with Latency-Associated Nuclear Antigen (LANA) to generate stable nuclear bodies (NBs). The LANA-associated nuclear bodies (LANA-NBs) are dependent on LLPS to develop dynamic structures that vary during the viral life cycle [60]. The LLPS of EBV proteins play a critical role in controlling the expression of host genes. EBV proteins EBNA2 and EBNALP regulate virus and cellular gene transcription as transcription factors, which can form liquid-like BMCs at Runx3 and MYC super-enhancer sites [61]. LLPS of EBNA2 mediates alternative RNA splicing patterns in cancer. EBNA2 modulates the aberrant splicing of MPPE1 through the recruitment of SRSF7 and SRSF1 to its motif [62]. In cells, the human papillomavirus (HPV) E2-p53 interaction involved in viral replication occurs via a direct interaction of proteins that promote heterotypic LLPS [63]. This interaction occurs specifically between the E2 DNA binding domain and the N-terminal transactivation domain of p53.

Cancer metabolism

LLPS has been identified as a mechanism that regulates the activity of enzymes within cells. Under hypoxia and high energy demands, glycolysis is elevated to stimulate cell proliferation. In hepatocarcinoma cells, hypoxia causes glycolytic enzymes such as the phosphofructokinase subunit Pfk2p to concentrate within membrane-less granules known as the glycolytic (G) body [64]. Snf1p, an AMP-activated protein kinase, is required for the formation of G bodies. G-body formation through LLPS is correlated with enhanced cell survival and proliferation in hypoxic stress conditions. Under hypoxic conditions, RNA is co-localized with glycolytic enzymes to the G bodies [65]. RNA scaffolding and recruitment are required for G body formation. Glycolysis is accelerated by the formation of G bodies through multivalent protein-RNA and protein–protein interactions. Under hypoxia, RNA facilitates the LLPS of glycolysis enzymes into G bodies. Condensed glycogen undergoes LLPS in vivo and in vitro, which results in the Laforin-Mst1/2 complex assembly in the glycogen droplet and triggers Yap. Glycogenolysis enzyme-liver glycogen phosphorylase (PYGL) or glucose-6-phosphatase (G6PC) deficiency results in glycogen storage diseases associated with Yap activation, and tumorigenesis [66].

LLPS in cancer therapy

With the advancement in knowledge of LLPS biology in cancer, the potential for developing effective cancer therapeutics has been raised, despite the challenges and obstacles that remain. There is potential therapeutic benefit from disrupting condensate formation through IDRs or physicochemical properties. Additionally, it has been proposed that selective modulation of LLPS, such as PTMs, may be an alternative therapeutic approach [18].
BMCs may form due to the thermodynamic instability of IDRs [6]. However, few IDR-binding compounds have been identified, and little attention has been paid to their ability to destabilize BMCs. Small molecule drugs are concentrated in BMCs, causing alterations in their on-target efficacy and pharmacodynamics [67]. There is increasing evidence that small molecules can bind to the IDRs of transcription factors such as TAF2, MYC, c-FOS, p53, and EWS and inhibit malignant cell transformation [13]. p53 structural mutants cause global destabilization, misfolding, and aggregation of the p53 protein. p53 aggregation can occur through a condensate-like state, and p53 can be found in protein BMCs. Small molecule compounds that interact with the p53 protein can be identified and used to dissolve the BMCs formed by the p53 structural mutant [52, 68]. Unlike p53 aggregation inhibitors, these identified small molecule compounds are efficacious against p53 BMCs and do not cause the reactivation of mutant p53 [68]. Several drugs, such as cisplatin, mitoxantrone, THZ1 (a CDK7 inhibitor), and tamoxifen were identified to selectively partition into BMCs formed by MED1. Mitoxantrone was also accumulated in nucleolar protein BMCs produced by NPM1 and FIB1 [67]. The BRD4 inhibitor was enriched in ED1, BRD4, and NPM1 BMCs [6]. The BET inhibitor that specifically targets BRD4 can operate by decreasing its LLPS [69]. Through the recruitment of BRD4 to the E3 ubiquitin ligase cereblon, ARV-825 (a BRD4 degrader) can induce efficient and sustained degradation of BRD4 in Burkitt’s lymphoma cell lines. ARV-825, a hetero-bifunctional proteolysis-targeting chimera (PROTAC), exerts a more effective inhibition of c-MYC levels and downstream signaling than small-molecule BRD4 inhibitors in Burkitt’s lymphoma cell lines, resulting in a significant decrease in cell proliferation and an increase in apoptosis (Fig. 3) [70, 71]. Chloroquine inhibits autophagy by impairing the autophagosome-lysosome fusion process, thereby increasing target protein levels (Fig. 4). Moreover, chloroquine prevents SQSTM1-positive autophagosomes from fusing with lysosomes, resulting in the buildup of STX17 [72].
Posttranslational modifications (PTMs) including arginine methylation, phosphorylation, arginine citrullination, ubiquitination, acetylation, and poly(ADP-ribosyl)ation mediate the condensation process and impact the properties of membrane-less compartments [73]. PTMs of RBPs as significant regulators of LLPS and RNP granule dynamics can decrease or strengthen the multivalent interactions between macromolecules as well as either exclude or recruit specific macromolecules from or into BMCs [74]. In the LLPS and RNP granule dynamics, arginine-methylation and phosphorylation function as important PTMs [75]. Assembling and disassembling SGs are regulated by deubiquitylases and small ubiquitin-like proteins, such as NEDD8 and SUMO [76]. Ubiquitin and ubiquitin-like proteins bind to target proteins covalently, resulting in a modified binding surface and protein interactions. Ubiquitin interacts with the UBA domain of ubiquitin 2 (UBQLN2), causing UBQLN2 BMCs to disperse [74].
Autophagy-initiated Atg1 complexes generate droplets via LLPS. Point mutations or phosphorylations that impede LLPS reduce the formation of pre-autophagosomal structure (PAS), a liquid-like condensate of Atg proteins [77]. Active TORC1 phosphorylates autophagy proteins, including Atg13, under high nutrient supply, disrupting the Atg13–Atg17 interaction and preventing the formation of PAS. In the absence of nutrients, TORC1 dephosphorylates autophagy proteins, including Atg13, which aggregates Atg1 molecules and autophosphorylates them. By activating Atg1, Atg13 can be re-phosphorylated, and PP2C phosphatases can reverse this process, thereby contributing to PAS condensate stability [78]. PAR polymerase (PARP) can regulate PARylation, and poly(ADP-ribose) glycohydrolase (PARG) hydrolyzes PAR to separate damaged DNA compartments. Following PARP1 activation, FUS is directed to sites of DNA damage where it forms the PARG reversible compartment. PARP1 inhibition hinders the formation of DNA damage repair foci and impedes the DNA damage repair process [79].
Various antitumor agents can form BMCs through physicochemical interactions, resulting in therapeutic benefits or drug resistance. The concentration and pharmacodynamics of drugs are affected by condensate characteristics. Antitumor agent cisplatin selectively concentrates in BMCs, promoting disease therapeutic progress [67, 80]. Tamoxifen, a common drug used to treat estrogen receptor (ER)-positive breast cancer, causes the expulsion of ERα from the MED1 BMCs [81, 82]. The ERα concentrates into the BMCs of MED1 in breast cancer cells in a tamoxifen-dependent, estrogen-dependent manner [67].
Some drugs that target transcription factors, hormone receptors, and nucleotide-binding proteins have the potential to prevent disease by inhibiting the formation of BMGs [18, 67]. Due to the heterogeneous and dynamic nature of IDR conformation, administering drugs against them can be challenging. Recently, new developments have been made in drugs that target IDRs. EPI-001 inhibits the androgen receptor by binding to its transduction unit 5 region (Tau-5) thereby delaying the onset of castration-resistant prostate cancer [83]. By controlling LLPS of the prion N-terminal domain in cancer, melatonin can suppress conformational modifications that can lead to aggregation, thus improving multidrug resistance [84].

Conclusion

Recent studies have demonstrated that BMCs play an important role in cellular processes. BMCs, formed by LLPS and membrane-free organelles, underlie biological processes and reactions. LLPS between macromolecules occurs as a result of multivalent interactions between macromolecules through modular domains, IDRs, and nucleic acid chains. LLPS is crucial in signal transduction, chromosomal disorganization, and transcriptional dysregulation. LLPS can be disrupted by gene mutations or epigenetic variants. LLPS can be governed by determining macromolecule concentration, targeting the components, or interfering with PTMs. Cancer-related molecules in BMCs are identified, and their modulation by LLPS may affect their function and progression. The mechanism responsible for the dynamic properties of LLPS in tumorigenesis and progression is not well understood. Studying LLPS in cancer progression will aid us in understanding the complex pathological processes of cancer and provide new treatment options.

Acknowledgements

Not applicable.

Declarations

Not applicable.
Not applicable.

Competing interests

The authors have no conflict of interest.
Open AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​. The Creative Commons Public Domain Dedication waiver (http://​creativecommons.​org/​publicdomain/​zero/​1.​0/​) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Literatur
1.
2.
3.
Zurück zum Zitat Banani SF, Lee HO, Hyman AA, Rosen MK. Biomolecular condensates: organizers of cellular biochemistry. Nat Rev Mol Cell Biol. 2017;18(5):285–98.PubMedPubMedCentralCrossRef Banani SF, Lee HO, Hyman AA, Rosen MK. Biomolecular condensates: organizers of cellular biochemistry. Nat Rev Mol Cell Biol. 2017;18(5):285–98.PubMedPubMedCentralCrossRef
4.
5.
Zurück zum Zitat Alberti S, Gladfelter A, Mittag T. Considerations and challenges in studying liquid–liquid phase separation and biomolecular condensates. Cell. 2019;176(3):419–34.PubMedPubMedCentralCrossRef Alberti S, Gladfelter A, Mittag T. Considerations and challenges in studying liquid–liquid phase separation and biomolecular condensates. Cell. 2019;176(3):419–34.PubMedPubMedCentralCrossRef
6.
Zurück zum Zitat Tong X, Tang R, Xu J, Wang W, Zhao Y, Yu X, Shi S. Liquid–liquid phase separation in tumor biology. Signal Transduct Target Ther. 2022;7(1):221.PubMedPubMedCentralCrossRef Tong X, Tang R, Xu J, Wang W, Zhao Y, Yu X, Shi S. Liquid–liquid phase separation in tumor biology. Signal Transduct Target Ther. 2022;7(1):221.PubMedPubMedCentralCrossRef
7.
Zurück zum Zitat Gomes E, Shorter J. The molecular language of membraneless organelles. J Biol Chem. 2019;294(18):7115–27.PubMedCrossRef Gomes E, Shorter J. The molecular language of membraneless organelles. J Biol Chem. 2019;294(18):7115–27.PubMedCrossRef
8.
Zurück zum Zitat Hong K, Song D, Jung Y. Behavior control of membrane-less protein liquid condensates with metal ion-induced phase separation. Nat Commun. 2020;11(1):5554.PubMedPubMedCentralCrossRef Hong K, Song D, Jung Y. Behavior control of membrane-less protein liquid condensates with metal ion-induced phase separation. Nat Commun. 2020;11(1):5554.PubMedPubMedCentralCrossRef
9.
Zurück zum Zitat Gao Y, Li X, Li P, Lin Y. A brief guideline for studies of phase-separated biomolecular condensates. Nat Chem Biol. 2022;18:1307–18.PubMedCrossRef Gao Y, Li X, Li P, Lin Y. A brief guideline for studies of phase-separated biomolecular condensates. Nat Chem Biol. 2022;18:1307–18.PubMedCrossRef
10.
11.
Zurück zum Zitat Feng Z, Chen X, Wu X, Zhang M. Formation of biological condensates via phase separation: characteristics, analytical methods, and physiological implications. J Biol Chem. 2019;294(40):14823–35.PubMedPubMedCentralCrossRef Feng Z, Chen X, Wu X, Zhang M. Formation of biological condensates via phase separation: characteristics, analytical methods, and physiological implications. J Biol Chem. 2019;294(40):14823–35.PubMedPubMedCentralCrossRef
12.
Zurück zum Zitat Bergeron-Sandoval LP, Safaee N, Michnick SW. Mechanisms and consequences of macromolecular phase separation. Cell. 2016;165(5):1067–79.PubMedCrossRef Bergeron-Sandoval LP, Safaee N, Michnick SW. Mechanisms and consequences of macromolecular phase separation. Cell. 2016;165(5):1067–79.PubMedCrossRef
13.
Zurück zum Zitat Ren J, Zhang Z, Zong Z, Zhang L, Zhou F. Emerging implications of phase separation in cancer. Adv Sci (Weinh). 2022;9(31):e2202855.PubMedCrossRef Ren J, Zhang Z, Zong Z, Zhang L, Zhou F. Emerging implications of phase separation in cancer. Adv Sci (Weinh). 2022;9(31):e2202855.PubMedCrossRef
14.
Zurück zum Zitat Guillen-Chable F, Bayona A, Rodriguez-Zapata LC, Castano E. Phase separation of intrinsically disordered nucleolar proteins relate to localization and function. Int J Mol Sci. 2021;22(23):13095.PubMedPubMedCentralCrossRef Guillen-Chable F, Bayona A, Rodriguez-Zapata LC, Castano E. Phase separation of intrinsically disordered nucleolar proteins relate to localization and function. Int J Mol Sci. 2021;22(23):13095.PubMedPubMedCentralCrossRef
15.
Zurück zum Zitat Li J, Zhang M, Ma W, Yang B, Lu H, Zhou F, Zhang L. Post-translational modifications in liquid–liquid phase separation: a comprehensive review. Mol Biomed. 2022;3(1):13.PubMedPubMedCentralCrossRef Li J, Zhang M, Ma W, Yang B, Lu H, Zhou F, Zhang L. Post-translational modifications in liquid–liquid phase separation: a comprehensive review. Mol Biomed. 2022;3(1):13.PubMedPubMedCentralCrossRef
16.
Zurück zum Zitat van der Lee R, Buljan M, Lang B, Weatheritt RJ, Daughdrill GW, Dunker AK, Fuxreiter M, Gough J, Gsponer J, Jones DT, et al. Classification of intrinsically disordered regions and proteins. Chem Rev. 2014;114(13):6589–631.PubMedPubMedCentralCrossRef van der Lee R, Buljan M, Lang B, Weatheritt RJ, Daughdrill GW, Dunker AK, Fuxreiter M, Gough J, Gsponer J, Jones DT, et al. Classification of intrinsically disordered regions and proteins. Chem Rev. 2014;114(13):6589–631.PubMedPubMedCentralCrossRef
17.
Zurück zum Zitat Wang W, Chen Y, Xu A, Cai M, Cao J, Zhu H, Yang B, Shao X, Ying M, He Q. Protein phase separation: a novel therapy for cancer? Br J Pharmacol. 2020;177(22):5008–30.PubMedPubMedCentralCrossRef Wang W, Chen Y, Xu A, Cai M, Cao J, Zhu H, Yang B, Shao X, Ying M, He Q. Protein phase separation: a novel therapy for cancer? Br J Pharmacol. 2020;177(22):5008–30.PubMedPubMedCentralCrossRef
18.
Zurück zum Zitat Mitrea DM, Mittasch M, Gomes BF, Klein IA, Murcko MA. Modulating biomolecular condensates: a novel approach to drug discovery. Nat Rev Drug Discov. 2022;21(11):841–62.PubMedPubMedCentralCrossRef Mitrea DM, Mittasch M, Gomes BF, Klein IA, Murcko MA. Modulating biomolecular condensates: a novel approach to drug discovery. Nat Rev Drug Discov. 2022;21(11):841–62.PubMedPubMedCentralCrossRef
19.
Zurück zum Zitat Lu J, Qian J, Xu Z, Yin S, Zhou L, Zheng S, Zhang W. Emerging roles of liquid–liquid phase separation in cancer: from protein aggregation to Immune-associated signaling. Front Cell Dev Biol. 2021;9:631486.PubMedPubMedCentralCrossRef Lu J, Qian J, Xu Z, Yin S, Zhou L, Zheng S, Zhang W. Emerging roles of liquid–liquid phase separation in cancer: from protein aggregation to Immune-associated signaling. Front Cell Dev Biol. 2021;9:631486.PubMedPubMedCentralCrossRef
20.
Zurück zum Zitat Hanahan D, Weinberg RA. Hallmarks of cancer: the next generation. Cell. 2011;144(5):646–74.PubMedCrossRef Hanahan D, Weinberg RA. Hallmarks of cancer: the next generation. Cell. 2011;144(5):646–74.PubMedCrossRef
21.
22.
Zurück zum Zitat Peng PH, Hsu KW, Wu KJ. Liquid–liquid phase separation (LLPS) in cellular physiology and tumor biology. Am J Cancer Res. 2021;11(8):3766–76.PubMedPubMedCentral Peng PH, Hsu KW, Wu KJ. Liquid–liquid phase separation (LLPS) in cellular physiology and tumor biology. Am J Cancer Res. 2021;11(8):3766–76.PubMedPubMedCentral
23.
Zurück zum Zitat Xiao Q, McAtee CK, Su X. Phase separation in immune signalling. Nat Rev Immunol. 2022;22(3):188–99.PubMedCrossRef Xiao Q, McAtee CK, Su X. Phase separation in immune signalling. Nat Rev Immunol. 2022;22(3):188–99.PubMedCrossRef
24.
Zurück zum Zitat Meng F, Yu Z, Zhang D, Chen S, Guan H, Zhou R, Wu Q, Zhang Q, Liu S, Venkat Ramani MK, et al. Induced phase separation of mutant NF2 imprisons the cGAS-STING machinery to abrogate antitumor immunity. Mol Cell. 2021;81(20):4147-4164.e7.PubMedCrossRef Meng F, Yu Z, Zhang D, Chen S, Guan H, Zhou R, Wu Q, Zhang Q, Liu S, Venkat Ramani MK, et al. Induced phase separation of mutant NF2 imprisons the cGAS-STING machinery to abrogate antitumor immunity. Mol Cell. 2021;81(20):4147-4164.e7.PubMedCrossRef
25.
Zurück zum Zitat Yu X, Zhang L, Shen J, Zhai Y, Jiang Q, Yi M, Deng X, Ruan Z, Fang R, Chen Z, et al. The STING phase-separator suppresses innate immune signalling. Nat Cell Biol. 2021;23(4):330–40.PubMedCrossRef Yu X, Zhang L, Shen J, Zhai Y, Jiang Q, Yi M, Deng X, Ruan Z, Fang R, Chen Z, et al. The STING phase-separator suppresses innate immune signalling. Nat Cell Biol. 2021;23(4):330–40.PubMedCrossRef
27.
Zurück zum Zitat Lin CW, Nocka LM, Stinger BL, DeGrandchamp JB, Lew LJN, Alvarez S, Phan HT, Kondo Y, Kuriyan J, Groves JT. A two-component protein condensate of the EGFR cytoplasmic tail and Grb2 regulates ras activation by SOS at the membrane. Proc Natl Acad Sci USA. 2022;119(19):e2122531119.PubMedPubMedCentralCrossRef Lin CW, Nocka LM, Stinger BL, DeGrandchamp JB, Lew LJN, Alvarez S, Phan HT, Kondo Y, Kuriyan J, Groves JT. A two-component protein condensate of the EGFR cytoplasmic tail and Grb2 regulates ras activation by SOS at the membrane. Proc Natl Acad Sci USA. 2022;119(19):e2122531119.PubMedPubMedCentralCrossRef
29.
Zurück zum Zitat Sampson J, Richards MW, Choi J, Fry AM, Bayliss R. Phase-separated foci of EML4-ALK facilitate signalling and depend upon an active kinase conformation. EMBO Rep. 2021;22(12):e53693.PubMedPubMedCentralCrossRef Sampson J, Richards MW, Choi J, Fry AM, Bayliss R. Phase-separated foci of EML4-ALK facilitate signalling and depend upon an active kinase conformation. EMBO Rep. 2021;22(12):e53693.PubMedPubMedCentralCrossRef
30.
Zurück zum Zitat Zhu G, Xie J, Kong W, Xie J, Li Y, Du L, Zheng Q, Sun L, Guan M, Li H, et al. Phase separation of disease-associated SHP2 mutants underlies MAPK hyperactivation. Cell. 2020;183(2):490-502.e18.PubMedPubMedCentralCrossRef Zhu G, Xie J, Kong W, Xie J, Li Y, Du L, Zheng Q, Sun L, Guan M, Li H, et al. Phase separation of disease-associated SHP2 mutants underlies MAPK hyperactivation. Cell. 2020;183(2):490-502.e18.PubMedPubMedCentralCrossRef
31.
Zurück zum Zitat Zhang JZ, Lu TW, Stolerman LM, Tenner B, Yang JR, Zhang JF, Falcke M, Rangamani P, Taylor SS, Mehta S, et al. Phase separation of a PKA regulatory subunit controls cAMP compartmentation and oncogenic signaling. Cell. 2020;182(6):1531-1544.e15.PubMedPubMedCentralCrossRef Zhang JZ, Lu TW, Stolerman LM, Tenner B, Yang JR, Zhang JF, Falcke M, Rangamani P, Taylor SS, Mehta S, et al. Phase separation of a PKA regulatory subunit controls cAMP compartmentation and oncogenic signaling. Cell. 2020;182(6):1531-1544.e15.PubMedPubMedCentralCrossRef
32.
Zurück zum Zitat Zaccolo M. cAMP buffering via liquid–liquid phase separation. Function (Oxf). 2021;2(1):zqaa048.PubMedCrossRef Zaccolo M. cAMP buffering via liquid–liquid phase separation. Function (Oxf). 2021;2(1):zqaa048.PubMedCrossRef
33.
Zurück zum Zitat Peng Q, Tan S, Xia L, Wu N, Oyang L, Tang Y, Su M, Luo X, Wang Y, Sheng X, et al. Phase separation in cancer: from the impacts and mechanisms to treatment potentials. Int J Biol Sci. 2022;18(13):5103–22.PubMedPubMedCentralCrossRef Peng Q, Tan S, Xia L, Wu N, Oyang L, Tang Y, Su M, Luo X, Wang Y, Sheng X, et al. Phase separation in cancer: from the impacts and mechanisms to treatment potentials. Int J Biol Sci. 2022;18(13):5103–22.PubMedPubMedCentralCrossRef
34.
Zurück zum Zitat Esposito M, Fang C, Cook KC, Park N, Wei Y, Spadazzi C, Bracha D, Gunaratna RT, Laevsky G, DeCoste CJ, et al. TGF-beta-induced DACT1 biomolecular condensates repress wnt signalling to promote bone metastasis. Nat Cell Biol. 2021;23(3):257–67.PubMedPubMedCentralCrossRef Esposito M, Fang C, Cook KC, Park N, Wei Y, Spadazzi C, Bracha D, Gunaratna RT, Laevsky G, DeCoste CJ, et al. TGF-beta-induced DACT1 biomolecular condensates repress wnt signalling to promote bone metastasis. Nat Cell Biol. 2021;23(3):257–67.PubMedPubMedCentralCrossRef
35.
Zurück zum Zitat Kang K, Shi Q, Wang X, Chen YG. Dishevelled phase separation promotes Wnt signalosome assembly and destruction complex disassembly. J Cell Biol. 2022;221(12):e202205069.PubMedCrossRef Kang K, Shi Q, Wang X, Chen YG. Dishevelled phase separation promotes Wnt signalosome assembly and destruction complex disassembly. J Cell Biol. 2022;221(12):e202205069.PubMedCrossRef
36.
Zurück zum Zitat Babcock RL, Pruitt K. Letting go: dishevelled phase separation recruits Axin to stabilize beta-catenin. J Cell Biol. 2022;221(12):e202211001.PubMedCrossRef Babcock RL, Pruitt K. Letting go: dishevelled phase separation recruits Axin to stabilize beta-catenin. J Cell Biol. 2022;221(12):e202211001.PubMedCrossRef
37.
Zurück zum Zitat Hofmann S, Kedersha N, Anderson P, Ivanov P. Molecular mechanisms of stress granule assembly and disassembly. Biochim Biophys Acta Mol Cell Res. 2021;1868(1):118876.PubMedCrossRef Hofmann S, Kedersha N, Anderson P, Ivanov P. Molecular mechanisms of stress granule assembly and disassembly. Biochim Biophys Acta Mol Cell Res. 2021;1868(1):118876.PubMedCrossRef
38.
Zurück zum Zitat Yang P, Mathieu C, Kolaitis RM, Zhang P, Messing J, Yurtsever U, Yang Z, Wu J, Li Y, Pan Q, et al. G3BP1 is a tunable switch that triggers phase separation to assemble stress granules. Cell. 2020;181(2):325-345.e28.PubMedPubMedCentralCrossRef Yang P, Mathieu C, Kolaitis RM, Zhang P, Messing J, Yurtsever U, Yang Z, Wu J, Li Y, Pan Q, et al. G3BP1 is a tunable switch that triggers phase separation to assemble stress granules. Cell. 2020;181(2):325-345.e28.PubMedPubMedCentralCrossRef
39.
Zurück zum Zitat Somasekharan SP, El-Naggar A, Leprivier G, Cheng H, Hajee S, Grunewald TG, Zhang F, Ng T, Delattre O, Evdokimova V, et al. YB-1 regulates stress granule formation and tumor progression by translationally activating G3BP1. J Cell Biol. 2015;208(7):913–29.PubMedPubMedCentralCrossRef Somasekharan SP, El-Naggar A, Leprivier G, Cheng H, Hajee S, Grunewald TG, Zhang F, Ng T, Delattre O, Evdokimova V, et al. YB-1 regulates stress granule formation and tumor progression by translationally activating G3BP1. J Cell Biol. 2015;208(7):913–29.PubMedPubMedCentralCrossRef
40.
Zurück zum Zitat Overwijn D, Hondele M. DEAD-box ATPases as regulators of biomolecular condensates and membrane-less organelles. Trends Biochem Sci. 2022;48:244–58.PubMedCrossRef Overwijn D, Hondele M. DEAD-box ATPases as regulators of biomolecular condensates and membrane-less organelles. Trends Biochem Sci. 2022;48:244–58.PubMedCrossRef
43.
Zurück zum Zitat Dao TP, Kolaitis RM, Kim HJ, O’Donovan K, Martyniak B, Colicino E, Hehnly H, Taylor JP, Castaneda CA. Ubiquitin modulates liquid–liquid phase separation of UBQLN2 via disruption of multivalent interactions. Mol Cell. 2018;69(6):965-978.e6.PubMedPubMedCentralCrossRef Dao TP, Kolaitis RM, Kim HJ, O’Donovan K, Martyniak B, Colicino E, Hehnly H, Taylor JP, Castaneda CA. Ubiquitin modulates liquid–liquid phase separation of UBQLN2 via disruption of multivalent interactions. Mol Cell. 2018;69(6):965-978.e6.PubMedPubMedCentralCrossRef
44.
45.
48.
Zurück zum Zitat Bouchard JJ, Otero JH, Scott DC, Szulc E, Martin EW, Sabri N, Granata D, Marzahn MR, Lindorff-Larsen K, Salvatella X, et al. Cancer mutations of the tumor suppressor SPOP disrupt the formation of active, phase-separated compartments. Mol Cell. 2018;72(1):19-36.e18.PubMedPubMedCentralCrossRef Bouchard JJ, Otero JH, Scott DC, Szulc E, Martin EW, Sabri N, Granata D, Marzahn MR, Lindorff-Larsen K, Salvatella X, et al. Cancer mutations of the tumor suppressor SPOP disrupt the formation of active, phase-separated compartments. Mol Cell. 2018;72(1):19-36.e18.PubMedPubMedCentralCrossRef
49.
Zurück zum Zitat Kamagata K, Kanbayashi S, Honda M, Itoh Y, Takahashi H, Kameda T, Nagatsugi F, Takahashi S. Liquid-like droplet formation by tumor suppressor p53 induced by multivalent electrostatic interactions between two disordered domains. Sci Rep. 2020;10(1):580.PubMedPubMedCentralCrossRef Kamagata K, Kanbayashi S, Honda M, Itoh Y, Takahashi H, Kameda T, Nagatsugi F, Takahashi S. Liquid-like droplet formation by tumor suppressor p53 induced by multivalent electrostatic interactions between two disordered domains. Sci Rep. 2020;10(1):580.PubMedPubMedCentralCrossRef
50.
Zurück zum Zitat Duffy MJ, Synnott NC, O’Grady S, Crown J. Targeting p53 for the treatment of cancer. Semin Cancer Biol. 2022;79:58–67.PubMedCrossRef Duffy MJ, Synnott NC, O’Grady S, Crown J. Targeting p53 for the treatment of cancer. Semin Cancer Biol. 2022;79:58–67.PubMedCrossRef
51.
Zurück zum Zitat Chen C, Fu G, Guo Q, Xue S, Luo SZ. Phase separation of p53 induced by its unstructured basic region and prevented by oncogenic mutations in tetramerization domain. Int J Biol Macromol. 2022;222(Pt A):207–16.PubMedCrossRef Chen C, Fu G, Guo Q, Xue S, Luo SZ. Phase separation of p53 induced by its unstructured basic region and prevented by oncogenic mutations in tetramerization domain. Int J Biol Macromol. 2022;222(Pt A):207–16.PubMedCrossRef
52.
Zurück zum Zitat Petronilho EC, Pedrote MM, Marques MA, Passos YM, Mota MF, Jakobus B, de Sousa GDS, Pereira da Costa F, Felix AL, Ferretti GDS, et al. Phase separation of p53 precedes aggregation and is affected by oncogenic mutations and ligands. Chem Sci. 2021;12(21):7334–49.PubMedPubMedCentralCrossRef Petronilho EC, Pedrote MM, Marques MA, Passos YM, Mota MF, Jakobus B, de Sousa GDS, Pereira da Costa F, Felix AL, Ferretti GDS, et al. Phase separation of p53 precedes aggregation and is affected by oncogenic mutations and ligands. Chem Sci. 2021;12(21):7334–49.PubMedPubMedCentralCrossRef
54.
56.
Zurück zum Zitat Altmeyer M, Neelsen KJ, Teloni F, Pozdnyakova I, Pellegrino S, Grofte M, Rask MD, Streicher W, Jungmichel S, Nielsen ML, et al. Liquid demixing of intrinsically disordered proteins is seeded by poly(ADP-ribose). Nat Commun. 2015;6:8088.PubMedCrossRef Altmeyer M, Neelsen KJ, Teloni F, Pozdnyakova I, Pellegrino S, Grofte M, Rask MD, Streicher W, Jungmichel S, Nielsen ML, et al. Liquid demixing of intrinsically disordered proteins is seeded by poly(ADP-ribose). Nat Commun. 2015;6:8088.PubMedCrossRef
57.
Zurück zum Zitat Kilic S, Lezaja A, Gatti M, Bianco E, Michelena J, Imhof R, Altmeyer M. Phase separation of 53BP1 determines liquid-like behavior of DNA repair compartments. EMBO J. 2019;38(16):e101379.PubMedPubMedCentralCrossRef Kilic S, Lezaja A, Gatti M, Bianco E, Michelena J, Imhof R, Altmeyer M. Phase separation of 53BP1 determines liquid-like behavior of DNA repair compartments. EMBO J. 2019;38(16):e101379.PubMedPubMedCentralCrossRef
58.
Zurück zum Zitat Zhang L, Geng X, Wang F, Tang J, Ichida Y, Sharma A, Jin S, Chen M, Tang M, Pozo FM, et al. 53BP1 regulates heterochromatin through liquid phase separation. Nat Commun. 2022;13(1):360.PubMedPubMedCentralCrossRef Zhang L, Geng X, Wang F, Tang J, Ichida Y, Sharma A, Jin S, Chen M, Tang M, Pozo FM, et al. 53BP1 regulates heterochromatin through liquid phase separation. Nat Commun. 2022;13(1):360.PubMedPubMedCentralCrossRef
59.
Zurück zum Zitat Fu H, Liu R, Jia Z, Li R, Zhu F, Zhu W, Shao Y, Jin Y, Xue Y, Huang J, et al. Poly(ADP-ribosylation) of P-TEFb by PARP1 disrupts phase separation to inhibit global transcription after DNA damage. Nat Cell Biol. 2022;24(4):513–25.PubMedPubMedCentralCrossRef Fu H, Liu R, Jia Z, Li R, Zhu F, Zhu W, Shao Y, Jin Y, Xue Y, Huang J, et al. Poly(ADP-ribosylation) of P-TEFb by PARP1 disrupts phase separation to inhibit global transcription after DNA damage. Nat Cell Biol. 2022;24(4):513–25.PubMedPubMedCentralCrossRef
60.
Zurück zum Zitat Vladimirova O, De Leo A, Deng Z, Wiedmer A, Hayden J, Lieberman PM. Phase separation and DAXX redistribution contribute to LANA nuclear body and KSHV genome dynamics during latency and reactivation. PLoS Pathog. 2021;17(1):e1009231.PubMedPubMedCentralCrossRef Vladimirova O, De Leo A, Deng Z, Wiedmer A, Hayden J, Lieberman PM. Phase separation and DAXX redistribution contribute to LANA nuclear body and KSHV genome dynamics during latency and reactivation. PLoS Pathog. 2021;17(1):e1009231.PubMedPubMedCentralCrossRef
61.
Zurück zum Zitat Peng Q, Wang L, Qin Z, Wang J, Zheng X, Wei L, Zhang X, Zhang X, Liu C, Li Z, et al. Phase separation of Epstein–Barr virus EBNA2 and its coactivator EBNALP controls gene expression. J Virol. 2020;94(7):e01771-19.PubMedPubMedCentralCrossRef Peng Q, Wang L, Qin Z, Wang J, Zheng X, Wei L, Zhang X, Zhang X, Liu C, Li Z, et al. Phase separation of Epstein–Barr virus EBNA2 and its coactivator EBNALP controls gene expression. J Virol. 2020;94(7):e01771-19.PubMedPubMedCentralCrossRef
62.
Zurück zum Zitat Peng Q, Wang L, Wang J, Liu C, Zheng X, Zhang X, Wei L, Li Z, Wu Y, Wen Y, et al. Epstein–Barr virus EBNA2 phase separation regulates cancer-associated alternative RNA splicing patterns. Clin Transl Med. 2021;11(8):e504.PubMedPubMedCentralCrossRef Peng Q, Wang L, Wang J, Liu C, Zheng X, Zhang X, Wei L, Li Z, Wu Y, Wen Y, et al. Epstein–Barr virus EBNA2 phase separation regulates cancer-associated alternative RNA splicing patterns. Clin Transl Med. 2021;11(8):e504.PubMedPubMedCentralCrossRef
64.
Zurück zum Zitat Jin M, Fuller GG, Han T, Yao Y, Alessi AF, Freeberg MA, Roach NP, Moresco JJ, Karnovsky A, Baba M, et al. Glycolytic enzymes coalesce in G bodies under hypoxic stress. Cell Rep. 2017;20(4):895–908.PubMedPubMedCentralCrossRef Jin M, Fuller GG, Han T, Yao Y, Alessi AF, Freeberg MA, Roach NP, Moresco JJ, Karnovsky A, Baba M, et al. Glycolytic enzymes coalesce in G bodies under hypoxic stress. Cell Rep. 2017;20(4):895–908.PubMedPubMedCentralCrossRef
65.
Zurück zum Zitat Fuller GG, Han T, Freeberg MA, Moresco JJ, Ghanbari Niaki A, Roach NP, Yates JR 3rd, Myong S, Kim JK. RNA promotes phase separation of glycolysis enzymes into yeast G bodies in hypoxia. Elife. 2020;9:e48480.PubMedPubMedCentralCrossRef Fuller GG, Han T, Freeberg MA, Moresco JJ, Ghanbari Niaki A, Roach NP, Yates JR 3rd, Myong S, Kim JK. RNA promotes phase separation of glycolysis enzymes into yeast G bodies in hypoxia. Elife. 2020;9:e48480.PubMedPubMedCentralCrossRef
66.
Zurück zum Zitat Liu Q, Li J, Zhang W, Xiao C, Zhang S, Nian C, Li J, Su D, Chen L, Zhao Q, et al. Glycogen accumulation and phase separation drives liver tumor initiation. Cell. 2021;184(22):5559-5576.e19.PubMedCrossRef Liu Q, Li J, Zhang W, Xiao C, Zhang S, Nian C, Li J, Su D, Chen L, Zhao Q, et al. Glycogen accumulation and phase separation drives liver tumor initiation. Cell. 2021;184(22):5559-5576.e19.PubMedCrossRef
67.
Zurück zum Zitat Klein IA, Boija A, Afeyan LK, Hawken SW, Fan M, Dall’Agnese A, Oksuz O, Henninger JE, Shrinivas K, Sabari BR, et al. Partitioning of cancer therapeutics in nuclear condensates. Science. 2020;368(6497):1386–92.PubMedPubMedCentralCrossRef Klein IA, Boija A, Afeyan LK, Hawken SW, Fan M, Dall’Agnese A, Oksuz O, Henninger JE, Shrinivas K, Sabari BR, et al. Partitioning of cancer therapeutics in nuclear condensates. Science. 2020;368(6497):1386–92.PubMedPubMedCentralCrossRef
68.
Zurück zum Zitat Lemos C, Schulze L, Weiske J, Meyer H, Braeuer N, Barak N, Eberspacher U, Werbeck N, Stresemann C, Lange M, et al. Identification of small molecules that modulate mutant p53 condensation. iScience. 2020;23(9):101517.PubMedPubMedCentralCrossRef Lemos C, Schulze L, Weiske J, Meyer H, Braeuer N, Barak N, Eberspacher U, Werbeck N, Stresemann C, Lange M, et al. Identification of small molecules that modulate mutant p53 condensation. iScience. 2020;23(9):101517.PubMedPubMedCentralCrossRef
69.
Zurück zum Zitat Wang C, Lu H, Liu X, Gao X, Tian W, Chen H, Xue Y, Zhou Q. A natural product targets BRD4 to inhibit phase separation and gene transcription. iScience. 2022;25(1):103719.PubMedCrossRef Wang C, Lu H, Liu X, Gao X, Tian W, Chen H, Xue Y, Zhou Q. A natural product targets BRD4 to inhibit phase separation and gene transcription. iScience. 2022;25(1):103719.PubMedCrossRef
70.
Zurück zum Zitat Lu J, Qian Y, Altieri M, Dong H, Wang J, Raina K, Hines J, Winkler JD, Crew AP, Coleman K, et al. Hijacking the E3 ubiquitin ligase cereblon to efficiently target BRD4. Chem Biol. 2015;22(6):755–63.PubMedPubMedCentralCrossRef Lu J, Qian Y, Altieri M, Dong H, Wang J, Raina K, Hines J, Winkler JD, Crew AP, Coleman K, et al. Hijacking the E3 ubiquitin ligase cereblon to efficiently target BRD4. Chem Biol. 2015;22(6):755–63.PubMedPubMedCentralCrossRef
71.
Zurück zum Zitat Wu S, Jiang Y, Hong Y, Chu X, Zhang Z, Tao Y, Fan Z, Bai Z, Li X, Chen Y, et al. BRD4 PROTAC degrader ARV-825 inhibits T-cell acute lymphoblastic leukemia by targeting ‘undruggable’ myc-pathway genes. Cancer Cell Int. 2021;21(1):230.PubMedPubMedCentralCrossRef Wu S, Jiang Y, Hong Y, Chu X, Zhang Z, Tao Y, Fan Z, Bai Z, Li X, Chen Y, et al. BRD4 PROTAC degrader ARV-825 inhibits T-cell acute lymphoblastic leukemia by targeting ‘undruggable’ myc-pathway genes. Cancer Cell Int. 2021;21(1):230.PubMedPubMedCentralCrossRef
72.
Zurück zum Zitat Mauthe M, Orhon I, Rocchi C, Zhou X, Luhr M, Hijlkema KJ, Coppes RP, Engedal N, Mari M, Reggiori F. Chloroquine inhibits autophagic flux by decreasing autophagosome-lysosome fusion. Autophagy. 2018;14(8):1435–55.PubMedPubMedCentralCrossRef Mauthe M, Orhon I, Rocchi C, Zhou X, Luhr M, Hijlkema KJ, Coppes RP, Engedal N, Mari M, Reggiori F. Chloroquine inhibits autophagic flux by decreasing autophagosome-lysosome fusion. Autophagy. 2018;14(8):1435–55.PubMedPubMedCentralCrossRef
73.
Zurück zum Zitat Luo YY, Wu JJ, Li YM. Regulation of liquid–liquid phase separation with focus on post-translational modifications. Chem Commun (Camb). 2021;57(98):13275–87.PubMedCrossRef Luo YY, Wu JJ, Li YM. Regulation of liquid–liquid phase separation with focus on post-translational modifications. Chem Commun (Camb). 2021;57(98):13275–87.PubMedCrossRef
74.
Zurück zum Zitat Hofweber M, Dormann D. Friend or foe-post-translational modifications as regulators of phase separation and RNP granule dynamics. J Biol Chem. 2019;294(18):7137–50.PubMedCrossRef Hofweber M, Dormann D. Friend or foe-post-translational modifications as regulators of phase separation and RNP granule dynamics. J Biol Chem. 2019;294(18):7137–50.PubMedCrossRef
75.
Zurück zum Zitat Giambruno R, Grzybowska EA, Fawzi NL, Dormann D. Editorial: the role of protein post-translational modifications in protein–RNA interactions and RNP assemblies. Front Mol Biosci. 2021;8:831810.PubMedCrossRef Giambruno R, Grzybowska EA, Fawzi NL, Dormann D. Editorial: the role of protein post-translational modifications in protein–RNA interactions and RNP assemblies. Front Mol Biosci. 2021;8:831810.PubMedCrossRef
76.
Zurück zum Zitat Jeon P, Ham HJ, Park S, Lee JA. Regulation of cellular ribonucleoprotein granules: from assembly to degradation via post-translational modification. Cells. 2022;11(13):2063.PubMedPubMedCentralCrossRef Jeon P, Ham HJ, Park S, Lee JA. Regulation of cellular ribonucleoprotein granules: from assembly to degradation via post-translational modification. Cells. 2022;11(13):2063.PubMedPubMedCentralCrossRef
77.
Zurück zum Zitat Fujioka Y, Alam JM, Noshiro D, Mouri K, Ando T, Okada Y, May AI, Knorr RL, Suzuki K, Ohsumi Y, et al. Phase separation organizes the site of autophagosome formation. Nature. 2020;578(7794):301–5.PubMedCrossRef Fujioka Y, Alam JM, Noshiro D, Mouri K, Ando T, Okada Y, May AI, Knorr RL, Suzuki K, Ohsumi Y, et al. Phase separation organizes the site of autophagosome formation. Nature. 2020;578(7794):301–5.PubMedCrossRef
79.
Zurück zum Zitat Singatulina AS, Hamon L, Sukhanova MV, Desforges B, Joshi V, Bouhss A, Lavrik OI, Pastre D. PARP-1 activation directs FUS to DNA damage sites to form PARG-reversible compartments enriched in damaged DNA. Cell Rep. 2019;27(6):1809-1821.e5.PubMedCrossRef Singatulina AS, Hamon L, Sukhanova MV, Desforges B, Joshi V, Bouhss A, Lavrik OI, Pastre D. PARP-1 activation directs FUS to DNA damage sites to form PARG-reversible compartments enriched in damaged DNA. Cell Rep. 2019;27(6):1809-1821.e5.PubMedCrossRef
81.
Zurück zum Zitat Leonard M, Zhang X. Estrogen receptor coactivator mediator subunit 1 (MED1) as a tissue-specific therapeutic target in breast cancer. J Zhejiang Univ Sci B. 2019;20(5):381–90.PubMedPubMedCentralCrossRef Leonard M, Zhang X. Estrogen receptor coactivator mediator subunit 1 (MED1) as a tissue-specific therapeutic target in breast cancer. J Zhejiang Univ Sci B. 2019;20(5):381–90.PubMedPubMedCentralCrossRef
82.
Zurück zum Zitat Cui J, Germer K, Wu T, Wang J, Luo J, Wang SC, Wang Q, Zhang X. Cross-talk between HER2 and MED1 regulates tamoxifen resistance of human breast cancer cells. Cancer Res. 2012;72(21):5625–34.PubMedPubMedCentralCrossRef Cui J, Germer K, Wu T, Wang J, Luo J, Wang SC, Wang Q, Zhang X. Cross-talk between HER2 and MED1 regulates tamoxifen resistance of human breast cancer cells. Cancer Res. 2012;72(21):5625–34.PubMedPubMedCentralCrossRef
83.
Zurück zum Zitat De Mol E, Fenwick RB, Phang CT, Buzon V, Szulc E, de la Fuente A, Escobedo A, Garcia J, Bertoncini CW, Estebanez-Perpina E, et al. EPI-001, a compound active against castration-resistant prostate cancer, targets transactivation unit 5 of the androgen receptor. ACS Chem Biol. 2016;11(9):2499–505.PubMedPubMedCentralCrossRef De Mol E, Fenwick RB, Phang CT, Buzon V, Szulc E, de la Fuente A, Escobedo A, Garcia J, Bertoncini CW, Estebanez-Perpina E, et al. EPI-001, a compound active against castration-resistant prostate cancer, targets transactivation unit 5 of the androgen receptor. ACS Chem Biol. 2016;11(9):2499–505.PubMedPubMedCentralCrossRef
Metadaten
Titel
Phase separation in cancer at a glance
verfasst von
Qingqing Xie
Jiejuan Cheng
Wuxuan Mei
Dexing Yang
Pengfei Zhang
Changchun Zeng
Publikationsdatum
01.12.2023
Verlag
BioMed Central
Erschienen in
Journal of Translational Medicine / Ausgabe 1/2023
Elektronische ISSN: 1479-5876
DOI
https://doi.org/10.1186/s12967-023-04082-x

Weitere Artikel der Ausgabe 1/2023

Journal of Translational Medicine 1/2023 Zur Ausgabe

Leitlinien kompakt für die Innere Medizin

Mit medbee Pocketcards sicher entscheiden.

Seit 2022 gehört die medbee GmbH zum Springer Medizin Verlag

RAS-Blocker bei Hyperkaliämie möglichst nicht sofort absetzen

14.05.2024 Hyperkaliämie Nachrichten

Bei ausgeprägter Nierenfunktionsstörung steigen unter der Einnahme von Renin-Angiotensin-System(RAS)-Hemmstoffen nicht selten die Serumkaliumspiegel. Was in diesem Fall zu tun ist, erklärte Prof. Jürgen Floege beim diesjährigen Allgemeinmedizin-Update-Seminar.

Gestationsdiabetes: In der zweiten Schwangerschaft folgenreicher als in der ersten

13.05.2024 Gestationsdiabetes Nachrichten

Das Risiko, nach einem Gestationsdiabetes einen Typ-2-Diabetes zu entwickeln, hängt nicht nur von der Zahl, sondern auch von der Reihenfolge der betroffenen Schwangerschaften ab.

Labor, CT-Anthropometrie zeigen Risiko für Pankreaskrebs

13.05.2024 Pankreaskarzinom Nachrichten

Gerade bei aggressiven Malignomen wie dem duktalen Adenokarzinom des Pankreas könnte Früherkennung die Therapiechancen verbessern. Noch jedoch klafft hier eine Lücke. Ein Studienteam hat einen Weg gesucht, sie zu schließen.

Battle of Experts: Sport vs. Spritze bei Adipositas und Typ-2-Diabetes

11.05.2024 DDG-Jahrestagung 2024 Kongressbericht

Im Battle of Experts traten zwei Experten auf dem Diabeteskongress gegeneinander an: Die eine vertrat die Auffassung „Sport statt Spritze“ bei Adipositas und Typ-2-Diabetes, der andere forderte „Spritze statt Sport!“ Am Ende waren sie sich aber einig: Die Kombination aus beidem erzielt die besten Ergebnisse.

Update Innere Medizin

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.