Abstract
Aggressive tumors pose ultra-challenges to drug resistance. Anti-cancer treatments are often unsuccessful, and single-cell technologies to rein drug resistance mechanisms are still fruitless. The National Cancer Institute defines aggressive cancers at the tissue level, describing them as those that spread rapidly, despite severe treatment. At the molecular, foundational level, the quantitative biophysics discipline defines aggressive cancers as harboring a large number of (overexpressed, or mutated) crucial signaling proteins in major proliferation pathways populating their active conformations, primed for their signal transduction roles. This comprehensive review explores highly aggressive cancers on the foundational and cell signaling levels, focusing on the differences between highly aggressive cancers and the more treatable ones. It showcases aggressive tumors as harboring massive, cancer-promoting, catalysis-primed oncogenic proteins, especially through certain overexpression scenarios, as predisposed aggressive tumor candidates. Our examples narrate strong activation of ERK1/2, and other oncogenic proteins, through malfunctioning chromatin and crosslinked signaling, and how they activate multiple proliferation pathways. They show the increased cancer heterogeneity, plasticity, and drug resistance. Our review formulates the principles underlying cancer aggressiveness on the molecular level, discusses scenarios, and describes drug regimen (single drugs and drug combinations) for PDAC, NSCLC, CRC, HCC, breast and prostate cancers, glioblastoma, neuroblastoma, and leukemia as examples. All show overexpression scenarios of master transcription factors, transcription factors with gene fusions, copy number alterations, dysregulation of the epigenetic codes and epithelial-to-mesenchymal transitions in aggressive tumors, as well as high mutation loads of vital upstream signaling regulators, such as EGFR, c-MET, and K-Ras, befitting these principles.
Similar content being viewed by others
Introduction
Aggressive cancers have been associated with multiple factors.1,2,3,4,5 This comprehensive review chronicles several highly aggressive cancers, their drug regimen, and their accelerated growth and malignancy. It explores the fundamental differences between highly aggressive cancers and the less aggressive, more treatable ones. Especially, it focuses on their underlying principles on the fundamental molecular and cell signaling levels. It formulates their hallmarks and weighs the clues that they provide to anti-cancer drug combination strategies targeting their ultra-strong drug resistance. Without exceptions, the examples of aggressive cancers that it narrates—all labeled by the clinical literature as highly aggressive—reveal scenarios promoting overexpression of oncogenic proteins, suppression of tumor suppressors, and crucially, altered expression of their regulators (e.g., see refs. 6,7,8,9,10,11; reviewed below). Potent oncogenic mutations, especially in upstream regulators, are common. The review further discusses the outcome: a massive payload of catalysis-ready conformational states, accompanied by powerful and heterogenous cell signaling.
Tumors harboring massive increases in the numbers of active molecules of relevant oncogenic proteins, especially through certain immense overexpression frameworks, are predisposed aggressive tumor candidates. Aggressive cancers are also manipulated by strong activating mutations in proteins upstream in major proliferation pathways, such as receptor tyrosine kinases (RTKs), constitutively contributing vast numbers of active molecules.12,13,14,15 Spatial single-cell transcriptomics16,17 support the expectation that over time, over- (for suppressors, under-) expression vandalizes the signaling networks, vacating cellular controls, amplifying dedifferentiation thus heterogeneity. Mutational variants upstream follow this pattern, as observed early on in knockin PIK3CA mutants,18,19,20,21 and RTKs such as epidermal growth factor receptor (EGFR), human epidermal growth factor receptor 2 (HER2, also known as ErbB2), and mesenchymal epithelial transition factor (MET, also known as hepatocyte growth factor receptor HGFR).22,23 The higher the number of the oncogenic proteins, the more undifferentiated the population, and the more aggressive the tumor.24,25,26 Undifferentiated, aggressive cancer states relate to the cargo of conformationally active, oncogenic molecules.27,28,29,30,31,32,33
Escalating active oncogenic (plummeting inactive suppressor) protein loads is a foundational hallmark of aggressive cancers (Fig. 1). It is “foundational” since it is expressed on the molecular level by conformational distributions,34,35,36 the most fundamental physical-chemical attribute of biomacromolecules. Dynamic conformational, or ensemble, propensities decide molecular, and cell function.37,38 Factors include timing (e.g., pediatric cancers have shorter time span to evolve, thus lower acquired mutation burden,26 with age there are more mutations39,40,41,42) and the function of the mutational variants (e.g., signaling, cell cycle). They also include predisposition through preexisting germline mutations,43,44,45,46,47,48 as in familial breast, ovarian, and colorectal cancers, and preexisting or acquired glioblastoma mutations as in the apparent linkage to earlier bouts of melanomas.49,50,51 Their tissue environment (e.g., brain, pancreas), and cell states are cardinal as well.52
The foundational principle underlying aggressive cancers: cancer aggressiveness by-the-numbers. We propose that the absolute number of active (oncogenic) conformations that the cancer harbors are a foundational hallmark of its aggressiveness. The higher the number—the more overspilled the signaling—the higher the heterogeneity. In aggressive cancers the number is extremely high. We dub this hallmark “cancer aggressiveness by-the-numbers”. In this molecular level definition, aggressive tumor candidates are those harboring massive, catalysis-primed oncogenic proteins, produced through overexpression scenarios and strong activating driver mutations. Both generate transcriptional landscapes signaling by-the-numbers scenario.14,25 Overexpression of oncogenic proteins is caused by an increase in gene expression due to epigenetic and genetic mechanisms, involving super-enhancers, hybrid gene fusions resulting from the combination of two independent genes, copy number alterations with lost or gained DNA segments, and an increase in signaling to target genes. A high propensity of active conformations shifts the population from an inactive state (gray spheres) to a constitutively active conformation (red RTKs and spheres).412 This leads to an increase in the number of active molecules (blue RTKs and spheres) of the corresponding protein node, resulting in an elevation in active transcription factors (small blue spheres with arrows), which in turn leads to the overexpression of oncogenes. Our molecular level definition updates the traditional definition of the National Cancer Institute, which defines an aggressive cancer as one that “forms, grows, or spreads rapidly and requires more intensive or severe treatment than usual”. For clarity, see the section on questions and clinical implications. It defines an active protein conformation, how their number can be assessed, explains why thresholds are still unavailable and challenging to establish. It also addresses how to measure expression of super-enhancers and overexpression, and more. Tumor suppressors are not included
Aggressive tumors are more difficult to treat, as they are made up of multiple states and can easily transition between them. For glioblastoma, four dominant lineage specific states were defined, neural-progenitor-like, oligodendrocyte-progenitor-like, astrocyte-like, and mesenchymal-like,53,54 but the states are nonhomogeneous with substates between them presenting a continuum. Consistently, analysis of the spectrum of tumor aggression between cells that are likely to remain in the primary tumor and those likely to metastasize, indicated differences in gene expression patterns.55 Aggressive clones turned on genes associated with epithelial-to-mesenchymal transition (EMT), showing that cells exist along a continuum of EMT states. Gene signatures of late-hybrid EMT states pointed to reduced survival in both human pancreatic and lung cancer patients, highlighting their relevance to clinical disease progression. Highly aggressive tumors are characterized by large increase of catalytically primed states.56 In glioblastoma, driver kinases include EGFR, platelet-derived growth factor receptor α (PDGFRA), and cyclin-dependent kinase 4 (CDK4). Their mechanistic framework may offer insight into therapeutic strategies to alleviate resistance to small molecule drug combinations.57,58,59,60,61,62,63 However, it has been unclear whether the states can be separately targeted, as the transition barriers are low, thus easy to flip. The lineage specific transcription factor waves, which reprogram neuroblastoma from self-renewal to differentiation,64 were also offered as targetable. The low barriers argue that the multitude of cell states in primary tumors would populate metastases of highly aggressive tumors too, challenging targeting.65,66,67,68,69,70
The absolute number of active (oncogenic) conformations that the cancer harbors can be a foundational hallmark of its aggressiveness (Fig. 1). The higher the number—the more the oncogenic signaling activates multiple pathways—the higher the heterogeneity. We call this molecular level hallmark “cancer aggressiveness by-the-numbers”. Below, we discuss highly aggressive tumor types and scenarios with altered transcriptomic landscapes, with examples. Lower grade tumors developing into highly aggressive forms harness these scenarios. Followed over time, spatial single-cell transcriptomics71 can help decipher mechanisms that tumor cells adopt toward therapeutic management.
Below, we focus on neuroblastoma, glioblastoma, pancreatic ductal adenocarcinoma (PDAC), non-small cell lung cancer (NSCLC), and leukemia. In the section on clinical research progress targeting the molecular mechanisms and targets of aggressive cancers, we add to these breast, liver, and colorectal cancers (CRC).
The dedifferentiated cancer state
Highly aggressive tumors are typically in undifferentiated or poorly differentiated states.72,73 They tend to grow rapidly and metastasize. The common postulate adopted by the National Cancer Institute is that the more abnormal the cells look under a microscope, the more aggressive the cancer and the faster it is likely to grow and spread. The differentiation stage of tumors is a key factor in histopathological classification of solid malignancies. A disorganized tumor under the microscope is more aggressive than a more differentiated tumor.74
The loss of organization and smooth borders in undifferentiated states are the outcome of corrupt protein-protein interaction networks which result from lopsided patterns of gene expression, normally regulated via homeostasis mechanisms. Grossly skewed over- (under-) expression of oncogenic genes, resulting from mutations and epigenetic alterations acquired during cancer evolution result in non-physiological patterns of genome expression, degrading the functional proteome.75,76,77,78,79 Normal developmental programs are temporally regulated by a network of interactions. The cells cascade down Waddington progenitor self-renewal growth states, exiting to accomplish complete differentiation.80,81,82,83 In contrast, rapidly proliferating mutant microclusters relapse into dedifferentiated populations,84,85,86 resembling earlier developmental states. In such lineage backsliding, cells do not revert to their original genome expression states. Instead, their dysregulated chromatin gains a high capacity to grow and proliferate. The cells are confined to Waddington wells, held captive by high energy mountains.
Function is the attribute of the differentiated state. Undifferentiated scenarios include amplified mutational load in key nodes, and rampant overexpression. As detailed in the examples below, certain overexpression scenarios appear predisposed to encode more aggressive tumors. Overexpression can be through super-enhancers, dysregulated epigenetic marks, especially associated with the transcription machinery, fusion of transcription factors, copy numbers of target genes, and signaling nodes, such as RTKs (Fig. 1). Aggressive, overexpressed pediatric tumors (e.g., PAX3-FOXO1 fusion-positive alveolar rhabdomyosarcoma) (Fig. 2), which are distinguished from the milder variant (embryonal rhabdomyosarcomas) provide a good example.87,88 Embryonal rhabdomyosarcomas do not harbor overexpression (under-expression) scenarios, nor does it display high mutation burden. Different than adult cancers, the relatively short developmental time span precludes high mutational burden populations. Instead, like other aggressive pediatric tumors, alveolar rhabdomyosarcoma harnesses an aberrant expression scenario, in this case an epigenetic mechanism, histone lysine demethylases. Epigenetically dysregulated super-enhancers in neuroblastoma and medulloblastoma provide additional examples.64,89,90
PAX3-FOXO1 fusion gene in rhabdomyosarcoma (RMS). RMS rare type of cancer that can be highly aggressive. It starts as a growth of cells in the soft tissue. There are several types of RMS, including embryonal (eRMS) and alveolar (aRMS), with aRMS being the far more aggressive. The PAX3-FOXO1 fusion gene is a signature genetic alteration for aRMS. It consists of a stable reciprocal translocation of chromosomes 2 and 13, t(2;13), which generates two derivative chromosomes, der(2) and der(13) (top left panel). The der(2) chromosome contains the FOXO1-PAK3 fusion gene, which encodes a protein lacking major functional domains, and the der(13) chromosome contains the PAX3-FOXO1 fusion gene, which encodes the PAX3-FOXO1 fusion transcription factor with enhanced transcriptional activity (top right panel). The PAX3 gene encodes a member of the PAX family of transcription factors. FOXO1 is also a transcription factor. In the PAX3-FOXO1 fusion, the in-frame DNA binding domain of PAX3 is fused with the transactivation domain of FOXO1, generating a transcription factor with powerful transcriptional power, altered post-translational regulation, and possibly new targets.88 As a pioneer factor, PAX3-FOXO1 alters the local chromatin structure and binding to repressed, inaccessible chromatin, and transcriptional activation (bottom panel). PAX3 paired box 3, FOXO1 forkhead box O1
Aberrant gene expression biases the cellular networks.91 Through crosslinked pathways, the oncogenic cargo corrupts multiple proliferation pathways, escalating self-renewal of nonfunctional, undifferentiated cell states, tumor progression, and treatment failure.54,92 Below, we discuss metastasis and follow with overexpression scenarios in aggressive cancers. We further suggest protocols for selection of drug combinations,57,58,59 and the underlying concept.
Aggressive cancers tend to metastasize
Apart from glioblastoma, the cancers described above undergo metastasis, a hallmark of aggressive cancer. Invasive growth sets the stage for metastatic dissemination.93 Invasion is a multi-step process. Cancer cells break away from a primary tumor mass and invade the surrounding stroma. Cascading metastasis events include diffusion, entering, and exiting dormancy, migration, and settling in distant organs.94 These capabilities can be gained by clonal selection from the heterogeneous genetic (e.g., mutational) states in the primary cancer cell population, and from transitioning states through epigenetic alterations and integration into the immune environment. Glioblastoma’s star-shaped astrocytes were offered as one reason for evading metastasis. The frequent metastasis of primary tumors to the brain, such as lung cancer, breast cancer and melanoma, were suggested to be the outcome of more productive pharmacology in those tumors as compare to glioblastoma.95,96,97,98 The longer elapsed time due to the effective drug regimen leads to emergence of additional resistant mutations, while the heterogeneously challenging blood-brain barrier compromises effective targeting of brain metastases. The process of metastasis raises fundamental questions, including a better grasp of the steps that are involved, their requirements, the proteins and cell states which are involved, and more, especially those related to pharmacology.98,99,100
Considering the connection between embryo development and cancer, proliferation and migration are not unique to metastasis. Embryonal brain development also involves proliferation of cells prior to acquiring fully differentiated cell states and migration. As in cancer, glial cells migrate individually or as groups,101,102,103 a process which is required for cortical development.102 In tumors, EMT cancer cells104 can spread as single cells via mesenchymal or amoeboid modes or move as groups, or microclusters.84,85,86,105 In tumors, cells clusters produce the growth factor epigen, observed in the nanolumina in cell-cell junction spaces, suggested to assist in migration, and as such, as therapeutic target.86 Circulating tumor cell clusters in breast cancer are up to 100 times more metastatic than single tumor cell, with the clustered cells displaying adhesion, epigenetic alterations and plasticity, pointing to high oncogenic and migration potential.106 Questions relating to how cells break away,104 how they seed secondary sites and how persist there, may relate to the embryonic brain too.
Oncogenic expression scenarios
Cell differentiation is controlled by dynamically balanced gene regulatory networks. The networks are influenced by lineage, state specific master transcription factors,64,107 which decide gene expression. While lineage populations over developmental time have not been quantified,108 distinct cell lineages likely accommodate nonidentical distributions, separated by Waddington energy barriers.81 Expression levels are determined by chromatin accessibility, which is governed by epigenetics. They are influenced by super-enhancers, which are clusters of enhancers whose activity is controlled by epigenetics marks.109 Temporal histone decorations decompress chromatin. Organized by scaffolding proteins, which can, e.g., catalyze the epigenetic marks of mono-, di-, or tri-methylation at lysine 4 of the peptide tails of histone H3, as a core component of the Set1/mixed-lineage leukemia (MLL) histone methyltransferase complexes,110 such as protein lysine methyltransferase G9a,111 result in spatially proximal, organized, master transcription factors. These bind to the super-enhancers and to other regulatory elements, driving stronger expression of their target genes. Cell type is manipulated by the combination of the active enhancers, thus master transcription factors.112 Exactly how the super-enhancer circuitry transitions the cell from a self-renewal state to a differentiated state is still unclear.64 But the resulting unrestricted overexpression spills over to connected pathways, activating them, and pushing the cell toward the dedifferentiated state—which is the core mechanism harnessed by aggressive cancers.113
Highly aggressive cancers include PDAC, NSCLC, CRC, hepatocellular carcinoma (HCC) or liver cancer, breast cancer, prostate cancer, and brain tumors, such as glioblastoma. Neuroblastoma is an example of a highly aggressive solid tumor in early childhood. Below we consider their oncogenic expression scenarios, starting with neuroblastoma.
Neuroblastoma
Neuroblastoma is a clinically heterogenous pediatric cancer. It can be aggressive, with diverse efforts and in-depth explorations under way.114,115,116,117,118,119,120 These include revealing its heterogeneity and clonal distributions,121,122,123 neuronal-like differentiation by downregulating CRMP5,124 and overexpression driving MYC-like gene expression.125 TAF1D transcriptionally activates G2/M phase-related genes in MYCN (a member of the MYC oncogene family encoding the transcription factor N-Myc), amplifying neuroblastoma,126 invasion and metastasis.127,128,129,130,131 Different from the differentiated phenotypic form,64,132,133,134 the aggressive, high-risk variant results from genetic and epigenic aberrations, driving high aberrant transcriptional output, dysregulated transcriptome, and the undifferentiated state.135 Neuroblastoma originates from the developing peripheral sympathetic nervous system64,136 and develops outside the brain and spinal cord—most commonly in or around the adrenal glands, on top of the kidneys.64 Embryonic neural crest cells proximal to the neural tube migrate and generate the ganglia of the peripheral sympathetic nervous system and the adrenal medulla.137 Genes that regulate the cell cycle (e.g., Liu et al.138) are overexpressed and commonly include MYCN,64,139 often with copy number alterations (CNAs). Elevated transcription promotes cell proliferation, stalling sympathoadrenal progenitor cells differentiating.140 N-Myc drives oncogenesis by cooperating with the G9a (also known as euchromatic histone-lysine N-methyltransferase 2, EHMT2) histone methyltransferase and the WD repeat-containing protein 5 (WDR5) adapter to mastermind global gene transcription.141,142 WDR5 assists N-Myc to bind promoters and up-regulate canonical Myc target genes to stimulate cell proliferation, whereas N-Myc recruits G9a to enhancers to down-regulate neuronal differentiation genes and inhibit cell differentiation.141 Crucial factors in aggressive neuroblastoma include core master transcription factor regulators of the neuroblastoma lineage PHOX2B, GATA3, and HAND2,64,143,144,145 with their cell type-specific overexpression determined by super-enhancers and N-Myc,131 whose levels in neuroblastoma migrating neural crest cells promote proliferation but restrain differentiation (Fig. 3).
Schematic diagram of N-Myc mediated gene overexpression in aggressive neuroblastoma. Neuroblastoma is a rare pediatric cancer that develops in the nervous system of infants and children. It affects immature nerve tissue (neuroblasts) in the adrenal glands. The MYCN gene is amplified in multiple neuronal and nonneuronal tumors.139 Among these, its amplification, which encodes the N-Myc transcription factor, is a key prognostic factor in neuroblastoma. N-Myc binds to available promoters containing the TATA box with the help of WDR5, a conserved regulator of gene expression.142 Binding upregulates canonical Myc target genes (e.g., ALK), activating and promoting RNA polymerase II (Pol II), driving oncogenic gene expression and cell proliferation. Transcription factors PHOX2B, HAND2 and GATA3 support N-Myc’s binding to the super-enhancers and promoter,131 and subsequent gene expression. Super-enhancers are formed by multiple enhancers. The gray area represents the mediator complex, which regulates transcription by connecting enhancers to promoters
Pediatric tumors are driven by relatively few genetic aberrations,135 carrying large genomic rearrangements and chromosomal CNAs, coupled with mutations in tumor suppressors or tumorigenic transcription factors, such as N-Myc. Considering their short evolution time, this is expected. In aggressive high-grade gliomas (pHGG), overexpression is controlled by histone methylation epigenetic alterations.146,147,148,149,150,151,152 In alveolar rhabdomyosarcoma, overexpression is via the PAX3-FOXO1 pioneer transcription factors gene fusion (Fig. 2).88,153 TRIB3 silencing promotes the downregulation of the AKT pathway and PAX3-FOXO1 in high-risk rhabdomyosarcoma.154
Among the remarkable innovative findings, is the potential of retinoic acid in neuroblastoma pharmacology.64 Retinoic acid, a derivative of vitamin A, can reprogram neuroblastoma cells and cause them to differentiate to neurons. This finding was recently highlighted at the National Cancer Institute.155 The most common combination of approved drugs against neuroblastoma includes cisplatin, carboplatin (Paraplatin), cyclophosphamide (Cytoxan), doxorubicin (Adriamycin), vincristine (Oncovin) and etoposide (Vepesid), but others may be used. For children in the high-risk group, other drugs might be added as well, and some drugs might be given at higher doses.156 Recently, a two hits rational dual strategy was proposed to more efficiently counter neuroblastoma,157 where a chimeric antigen receptor (CAR)-T cell treatment appears more efficient by increasing the cell surface expression of the CAR target structure via a small molecule.158,159,160 CAR-T cells targeting anaplastic lymphoma kinase (ALK), which is frequently highly expressed on the surface of neuroblastoma cells, eliminate the tumor cells.
Additional targeting explorations include TIM-3 blockade,161 a potential targetable mutation,162 targeting immune checkpoint,163 and its migration, invasion and metastasis,164 warning about the consequences of radiation,165,166 and chemotherapy, ALK-related neuroblastic tumor susceptibility, targeting c-Myc transactivation, immune modulation, prognosis, suppressor of ferroptosis, prevention, and more.167,168,169,170
Glioblastoma
Glioblastoma, the deadliest brain cancer, was proposed to be spatially organized by neurodevelopmental programs, mimic glial-like wound healing, and characterized by substantial heterogeneity.171,172 Its dynamic organization, including proliferating and differentiated cells, retains the hierarchy of normal brain development. Glioblastoma’s tumor cells epitomize neural cell types.54,173 Their relative frequencies are influenced by the copy number of CDK4, EGFR, and PDGFRA and mutations in NF1. Glioblastoma aggressiveness, and therapeutic failure, are rooted in its excessive transcriptional heterogeneity, resulting from overexpression and activating mutations in the proteins it targets in the cell cycle and proliferative signaling.54,174 Its variability is intra-tumoral, temporal, and influenced by therapy.175 The states are plastic, encompassing multiple transitional microstates, and the heterogeneity is impacted the tumor’s developmental state, which embraces early embryonic neural development. Glioblastoma’s mammoth transcriptomic heterogeneity emerges via specific core genetic events that corrupt cell signaling, epigenetic, developmental, and microenvironmental origins. Within this framework, the immense amplification of the mitogen-activated protein kinase (MAPK) pathway, including via EGFR and PDGFRA, plays a cardinal role. EGFR gene amplifications are associated with astrocyte-like (AC-like) cells, CDK4 amplification with neural-progenitor-like (NPC-like), and PDGFRA with oligodendrocyte-progenitor-like (OPC-like) states. Severe oncogenic phenotype in patients with large deletions of NF1 Chr5q (chromosome 5q) point to higher frequency of mesenchymal-like (MES-like) states, leading to increased plasticity, pro-metastatic traits such as increased motility, invasiveness, and immune system evasion, typically observed in the pro-neural subtype. Mesenchymal transformation has been dubbed the Rosetta stone of glioblastoma pathogenesis.176
We provide an overview of the signaling pathways whose aberration and overexpression contribute to glioblastoma aggressiveness along with some drugs targeting these pathways (Fig. 4). Overexpression resulting from strong amplification or activating mutations in RTKs, such as EGFR, PDGFRα, VEGFR, and ALK fusion protein with echinoderm microtubule-associated protein-like 4 (EML4), triggers overspilled signaling through pathway crosstalk, promoting enormous heterogeneity, driving dedifferentiation and drug resistance. Several therapeutic agents have been identified that target key components of these signaling pathways. For instance, EGFR is targeted by small molecule inhibitors such as osimertinib (Tagrisso), mavelertinib (PF-06747775), and naquotinib (ASP8273), while VEGFR is targeted by the monoclonal antibody bevacizumab (Avastin, Mvasi, Zirabev). ALK fusion proteins can be inhibited by agents like crizotinib (Xalkori), and ceritinib (Zykadia), alectinib (Alecensa), brigatinib (Alunbrig), and lorlatinib (Lorbrena). PI3K inhibitors include alpelisib (Piqray), RLY-2608, STX-478, and LOXO-783, while AKT inhibitors encompass GSK690693, GDC-0068, AZD5363, MK-2206, ARQ-092, and TAS-117. Additionally, Ras inhibitors such as tipifarnib (Zarnestra), sotorasib (Lumakras), and deltarasin, alongside the downstream MEK inhibitor trametinib (Mekinist), serve to attenuate oncogenic signaling. For chemotherapeutic agents specific to brain cancers, temozolomide (Temodar) prevents cancer cells from making DNA, leading to cell cycle arrest at G2/M and apoptosis, and carmustine (BiCNU, Gliadel) cross-links DNA and RNA, resulting in inhibition of DNA synthesis, RNA production and RNA translation.
Signaling pathways in glioblastoma. PI3K/AKT, MAPK, and JAK/STAT signaling pathways orchestrate cell growth, proliferation, survival, and apoptosis. Their dysregulation is a critical driver of glioblastoma aggressiveness. In the PI3K/AKT pathway, downstream targets like ribosomal proteins rpS6 and eIF4E foster cell growth through mTOR signaling. Loss-of-function mutations in the tumor suppressor NF1 abrogate its inhibitory role on Ras, leading to overactivation of the MAPK pathway. The MAPK and JAK/STAT pathways regulate cell proliferation and survival via the transcription factors such as c-Myc, Elk-1, c-Jun, and STAT3/5. VEGFR vascular endothelial growth factor receptor, EML4 echinoderm microtubule-associated protein-like 4, eIF4E eukaryotic translation initiation factor 4E
By corrupting cell signaling, overexpression of these RTKs alters the landscape of the cellular states, enforcing tumor development. Overexpression of EGFR in nestin+ neural progenitors drive astrocytoma-like tumors; PDGF/PDGFR manipulate oligodendroglioma-like tumors,54,177 activating primarily the Janus kinase/signal transducer and activator of transcription 3 (JAK/STAT3), phosphoinositide 3-kinase/protein kinase B/mammalian target of rapamycin (PI3K/AKT/mTOR), Ras/MAPK, and transforming growth factor β (TGFβ) pathways (Fig. 4). In glioblastoma, EPHA2, an RTK, was observed to mediate PDGFA/PDGFRA signaling, and EGFR, a major driver of MAPK, activates glutamate dehydrogenase (GDH1) transcription to promote glutamine metabolism through the mitogen-activated protein kinase kinase/extracellular signal-regulated kinase/ETS domain-containing protein Elk-1 (MEK/ERK/Elk-1) pathway.178 The outcome is biased signaling, skewed toward different dedifferentiated cell states, whose relative frequencies reflect the amplification genetic events. The multiple affected tumor pathways scenarios argue why combination therapy targeting multiple pathways is essential.
Currently, there is no small drug treatment approach that will be effective for every glioblastoma. Vorasidenib had positive results in delaying progression of a specific form of glioma.179,180 It doubled progression-free survival in people with recurrent grade 2 glioma with isocitrate dehydrogenase 1 (IDH1) or 2 (IDH2) mutations. Not a small molecule, but important to cite, is intrathecal bivalent CAR-T cells targeting EGFR and interleukin-13 receptor α2 (IL-13Rα2), which is undergoing clinical trials in recurrent glioblastoma.181 It is a dual-target treatment, with CAR-T cells targeting proteins common in brain tumors: EGFR (in 60% of the patients), and IL-13Rα2 (in 75%).182 A tested drug combination includes indotecan (LMP400, a topoisomerase I inhibitor) and niraparib (PARP inhibitor, prevents DNA repair, leading to cell death).183 Drugs used to treat glioblastoma multiforme184 include temozolomide (Temodar), bevacizumab (Avastin, Mvasi, Zirabev), procarbazine hydrochloride (Matulane), hydroxyurea (Hydroxycarbamide, Droxia, Siklos), carmustine (BiCNU, Gliadel), bevacizumab-maly (Alymsys), bevacizumab-tnjn (Avzivi), and bevacizumab-adcd (Vegzelma).
Additional glioblastoma mechanistic studies were carried out.185,186 Pharmacological explorations include degradation185 and WDR1-dependent cytoskeleton remodeling.187 They target its signaling,188,189,190 starve it, target its polymerase, and its epigenetic STING modulation.191 They also undertake prognosis,192 target the immune checkpoint,193,194 immunotherapy,195,196 its invasion via EMT, take up peptide-based inhibition strategy,197 microRNAs, chemotherapy, and more.198,199,200 Glioblastoma heterogeneity has been undertaken through network modeling and a systems-level approaches, which identify shared and tumor-specific signaling alterations, such as MEK1 activation, NUMB variability, and 3D mutation patches, ultimately stratifying patients into groups with distinct survival outcomes and advancing network-guided precision medicine.201,202 Several single-cell and spatial analysis studies have been aiming to map and analyze the spatial heterogeneity of glioblastoma to reveal tumor evolution, regional molecular specificity, and potential therapeutic targets.172,203,204,205
Pancreatic ductal adenocarcinoma (PDAC)
While the nature of PDAC cell of origin is unclear,206 early emergence of mutations in KRAS [G12D (39%), G12V (32%), G12R (17%)] in plastic pancreatic exocrine cells is a hallmark of PDAC.207 The mutations occur in early stages,208 initiating and maintaining PDAC development. High frequencies of strong KRAS mutations testify to potent, heterogenous signaling output through multiple effector pathways (Fig. 5). Further contributing to basal cell transitions and intrapopulation heterogeneity are the mutations in CDKN2A (a cyclin-dependent kinase inhibitor), TP53 (a tumor suppressor), and SMAD4 (a transcription factor mediating TGFβ signaling), all associated with intraepithelial neoplasia (PanIN), the dominant precursor of PDAC.209 Overexpression of the yes-associated protein 1 (YAP1) transcription factor210 and the frequently mutated epigenetic regulatory genes, including histone modification enzymes, overexpression of the histone deacetylase 5 (HDAC5),211 aberrations in MLL histone methylases, histone methyltransferases, and the lysine demethylase 6A (KDM6A) histone demethylase, a potent tumor repressor,212,213,214,215 which is associated with histone 3 lysine 4 (H3K4) methylation and histone 3 lysine 27 (H3K27) demethylation,216 aggressively increase the transitions and heterogeneity. Tumors with genetic defects in MLLs are likely to induce expression of chromatin-regulating genes and cell proliferation-associated genes (including SWI/SNF chromatin remodeling complexes, and their components) and genes involved in cell cycle progression and proliferation.215 HDAC1 and HDAC2 are also highly expressed in pancreatic cancer, and can be recruited to the epithelial-cadherin (CDH1) promoter (involved in histone deacetylation) by zinc-finger E-box binding homeobox 1 (ZEB1), thereby promoting EMT and tumor metastasis.217,218 SIRT6 (Sirtuin 6) histone deacetylase ablation can promote PDAC metastasis by hyperacetylating histone 3 lysine 9 (H3K9) and histone 3 lysine 56 (H3K56), leading to Myc recruitment. These events result in overexpression of high-mobility group AT-hook 2 (HMGA2), insulin-like growth factor 2 mRNA-binding proteins (IGF2BP1 and IGF2BP3), downstream of let-7.219 Overexpression of c-Myc is crucial hallmark of aggressive cancer cells.220,221 Recently it was also observed K-Ras-GTP inhibition in pancreatic cancer.222 The overexpression patterns of Myc, IGF2BP2, a tumor promoter that drives cancer proliferation through its client mRNAs IGF2 and HMGA1, which regulates the expression of replication-dependent histone genes and the cell-cycle in cancer cells, and epigenetics gene aberrations, couple with overactive signaling pathways in pancreatic cancer tumorigenesis and metastasis, including the MAPK, PI3K/AKT, NF-κB, JAK/STAT, Hippo/YAP, and Wnt pathways. Further, YAP, a transcriptional coactivator along with transcriptional co-activator with PDZ binding motif (TAZ) and transcriptional enhanced associate domain (TEAD) is also overexpressed.211,222,223,224
Schematic representation of four major molecular lesions in pancreatic ductal adenocarcinoma (PDAC). PDAC is primarily caused by genetic mutations in four genes: an oncogene, KRAS (encoding K-Ras); and three tumor suppressor genes, TP53 (encoding p53, a transcription factor), CDKN2A (encoding p16INK4A, a CDK inhibitor), and SMAD4 (encoding SMAD4, a transcription factor). Constitutively active K-Ras with the G12X mutation leads to increased activation of Raf, MEK, and ERK through a phosphorylation cascade. ERK activates transcription factors such as c-Myc, Elk-1, and c-Jun, leading to cell proliferation. Insulin binding to IR together with active K-Ras initiates PI3K activation. PI3K converts PIP2 to PIP3, leading to mTORC1 activation. This includes AKT activation by PDK1 and mTORC2. mTORC1 phosphorylates S6K1 and 4E-BP1. S6K1 activates rpS6. Phosphorylation of 4E-BP1 removes its inhibitory role on eIF4E, which is involved in translational activation and regulation of cell growth. Inactivation of p16INK4A by mutation or genomic deletion impairs its function as a CDK4 inhibitor, leading to an unregulated cell cycle transition. Inactivation of p53 by mutation hinders its functions, such as blocking of angiogenesis, DNA repair, and induction of apoptosis. p53 mutant also impairs the expression of p21 (a CDK inhibitor), which is involved in G1/S arrest due to damaged DNA. The TGFβ ligand binds to the TGFβ receptor type II (RII) dimer, which recruit the type I (RI) dimer to form a hetero-tetrameric complex. RII phosphorylates the serine/threonine kinase of RI. Under physiological conditions, RI phosphorylates the receptor-regulated SMAD (RSMAD), such as SMAD2 and SMAD3, causing them to dissociate from the receptor complex. The RSMAD complex associates with a common mediator SMAD (coSMAD), i.e., SMAD4, to form a complex that enters the nucleus to bind its target genes, leading to cell cycle arrest and apoptosis. Inactivation of SMAD4 by mutation or genomic deletion impairs its tumor suppressor function in PADC. As discussed in the text, as a highly aggressive cancer, it also involves overexpression of multiple genes, e.g., YAP1, MYC, HMGA2, IGF2BP1 and IGF2BP3, and dysregulation of epigenetics modulators, e.g., HDAC1/2/5, KDM6A, MLL histone methylases, and histone methyltransferases. IR insulin receptor, PDK1 phosphoinositide-dependent kinase 1, 4E-BP1 eukaryotic translation initiation factor 4E (eIF4E)-binding protein 1, IRS insulin receptor substrate
Through gains in copy numbers, overexpressed transcription factors and their activators upregulate transcription of multiple genes, thereby promoting EMT. Key signaling nodes like AKT have been pointed as harboring aberrant expression, as are epigenetics-related proteins, such as HDAC1 and HDAC2, MLL histone methylases, histone methyltransferases, and the KDM6A histone demethylase, and more. These combine with RAS, CDKN2A, TP53, and SMAD4 high oncogenic mutational burden (Fig. 5), collectively resulting in multiple distorted signaling pathways, pushing tumor cells dedifferentiation. Altogether, these stand out, clarifying the aggressive behavior of PDAC, which can be attributed to the multiple powerful overexpression scenarios, promoting EMT, and shifts in cell states.
Drug resistance to RMC-7977, is a highly selective inhibitor of active K-Ras, H-Ras, and N-Ras, resulted in detected Myc copy number gain, which could be overcome by combinatorial TEAD inhibition in vitro.222 c-Myc overexpression is common in PDAC. It binds to promoters, interacts with proliferative pathways in pancreatic cancer, boosts aerobic glycolysis and regulates glutamate biosynthesis, thus cancer cells metabolism, contributing to EMT, and aggressive PDAC behavior.220 Small drug therapies approved for pancreatic cancer225 include paclitaxel (Abraxane, Taxol), everolimus (Afinitor, Votubia, Zortress), capecitabine (Xeloda), erlotinib hydrochloride (Tarceva), 5-fluorouracil or 5-FU (Carac, Tolak, Efudex, Fluoroplex), gemcitabine hydrochloride (Gemzar, Infugem), irinotecan hydrochloride (Campto, Camptosar, Onivyde), olaparib (Lynparza), mitomycin (Mitosol, Mutamycin), and sunitinib malate (Sutent).
Additional exploratory therapeutic approaches target reprogramming of cancer-associated fibroblasts, heterogeneity, metastatic pancreatic cancer,226 degradation,227 neddylation,228 nutrient restriction, vascular endothelial growth factor (VEGF)-independent angiogenesis, targeting ferroptosis, signaling,229 metabolic dependencies,230 chemotherapy-induced anti-tumor immunity to boost anti-PD-1 therapy, modulating the immune response including immune checkpoint therapy, prognosis, EMT-associated chemoresistance, prevention, enhancing proteasome inhibition, IL-33, and more, altogether, broad translational challenges and trends.
Non-small cell lung cancer (NSCLC)
Lung cancer is a leading cause of cancer-related mortality.231 NSCLC is more common, especially lung adenocarcinoma (LUAD), and less aggressive than small cell lung cancer (SCLC). Both metastasize quickly, often to the brain.232,233,234,235 Its expression patterns were comprehensively analyzed,236 and its major signaling tracked to RTK, MAPK, PI3K/AKT/mTOR, JAK/STAT, apoptosis [B cell lymphoma protein, Bcl-2-associated X protein, first apoptosis signal ligand (FasL)], Notch, Hedgehog, Wnt, and the YAP/TAZ/TEAD Hippo pathways.237,238
NSCLC tumors commonly harbor overexpression of diverse catalytically primed kinases (Fig. 6). Scenarios include NTRK fusion genes, such as NTRK1, NTRK2, and NTRK3 (encoding tropomyosin receptor kinases, TrkA, TrkB, and TrkC, respectively),239,240 that link the tyrosine kinase domain with a 5’ fusion partner (over 50 partners across cancers). With no ligand binding domain, the resulting constitutively activated Trk promotes tumor cell proliferation. A second scenario includes 5´ BRAF fusion partners, the most frequent being AGK (a gene encoding acylglycerol kinase) in LUAD.241,242,243 Most are in-frame with the B-Raf kinase domain,242,243 with the N-terminal Ras binding domain (an autoinhibitory domain) truncated,244 making B-Raf and the MAPK pathway constitutively active. In a third scenario, NSCLC harnesses metastatic EGFR fusions,245 where the tyrosine kinase domain is frequently fused to RAD51, a protein involved in DNA damage response. The missing EGFR autophosphorylation and adapter binding sites at its C-terminal tail are replaced by EGFR’s Tyr845 in the EGFR kinase domain-RAD51 fused protein.245,246 The loss of tyrosine 1045, which acts in EGFR degradation, extends the protein half-life time. Additional kinase fusion scenarios include the LUAD ALK-positive fusion often with EML4.247,248 HER2 tyrosine kinase receptor also undergoes alterations, including amplification, mutations, and overexpression.249 Rearranged during transfection (RET) overexpression is frequent in lung neuroendocrine tumors and is associated with response to RET tyrosine kinase inhibitors.250 MET overexpression ranges in from 15 to 70% of NSCLC patients.251 MET oncogenic mechanisms include fusions, mutations in the tyrosine kinase domain, and exon 14 skipping alterations.252 c-MET is an RTK for hepatocyte growth factor (HGF). Like some other RTKs, c-MET activates multiple pathways, including MAPK and PI3K/AKT/mTOR. Activating mutations, such as MET exon 14 skipping mutations (METex14), cause hyper-activation, thus proliferation, EMT, and metastasis, especially to the brain,232,253 previously associated with a poor prognosis, currently targeted with capmatinib (Tabrecta).254
Hallmarks of lung adenocarcinoma. Overexpression of active conformations such as catalytically primed kinases is associated with non-small cell lung cancer (NSCLC). These increased kinase activities are attributed to fusion genes, including NTRK fusion, AGK-BRAF fusion, EGFR-RAD51 fusion, EML4-ALK fusion, and RET fusion genes (see text for details). NSCLC also harbors HER2 alterations, MET overexpression, and K-Ras G12C mutation. Generic names of small drugs that inhibit these active molecules are shown in the figure with the brand name in parentheses
All RTKs sustain residue mutations (recently reviewed by Waarts et al.255). In drug resistance the alterations can collaborate, as observed in the METex14 driver mutation co-existing with amplification of CDK4 (which inactivates retinoblastoma, a regulator of cell proliferation) and MDM2 (inactivates p53), PDGFR, and FGFR. Other possible alterations include EGFR, PDGFR, RET, and K-Ras. In individual patients a large majority of driver gene mutations are homogeneous across all metastases,256 but rare subclonal alterations are heterogeneous in different metastasis.232,257 These suggest that a single anti-resistance drug cannot slow tumor growth. However, combined with drugs targeting the dominant driver mutations (e.g., MET14ex), may extend progression-free and overall survival.232 Amplification of Myc and TERT (telomerase reverse transcriptase) appears common events in lung cancer patients. This is not surprising since Myc copy number change also emerges in drug resistance to K-Ras mutations, as shown recently in targeting K-RasG12C in pancreatic cancers.258
Small drug therapy for NSCLC includes cisplatin, carboplatin (Paraplatin), pemetrexed (Alimta), paclitaxel (Abraxane, Taxol), docetaxel (Docefrez, Taxotere), gemcitabine hydrochloride (Gemzar, Infugem), and vinorelbine (Navelbine).259 Angiogenesis inhibitors including bevacizumab (Avastin, Mvasi, Zirabev) are used in combination with chemotherapy, immunotherapy, or erlotinib (Tarceva) in advanced or metastatic NSCLC. Ramucirumab (Cyramza) is used in combination with the targeted drug erlotinib (Tarceva) or chemotherapy advanced or metastatic NSCLC. Specific targeting of oncogenic proteins includes K-Ras inhibitors, sotorasib (Lumakras) for advanced NSCLC with the K-Ras G12C mutation.260 Adagrasib (Krazati) resembles sotorasib. EGFR inhibitors targeting cells with either an exon 19 or exon 21 mutation include osimertinib (Tagrisso), afatinib (Gilotrif), erlotinib (Tarceva), dacomitinib (Vizimpro), gefitinib (Iressa), or erlotinib (Tarceva) in combination with a VEGF inhibitor. EGFR inhibitors that target cells with S768I, L861Q and/or G719X mutations include afatinib (Gilotrif) or osimertinib (Tagrisso), and erlotinib (Tarceva), dacomitinib (Vizimpro), and gefitinib (Iressa). EGFR inhibitors that target cells with an exon 20 mutation include amivantamab (Rybrevant) monoclonal antibody. ALK inhibitors include lorlatinib (Lorbrena) and second-generation ALK inhibitors alectinib (Alecensa), brigatinib (Alunbrig), and ceritinib (Zykadia). ROS1 inhibitors include entrectinib (Rozlytrek), crizotinib (Xalkori), and ceritinib (Zykadia). A B-Raf inhibitor, dabrafenib (Tafinlar), is used in combination with trametinib (Mekinist), a MEK inhibitor, and encorafenib (Braftovi) with binimetinib (Mektovi), also a MEK inhibitor. Vemurafenib (Zelboraf) and dabrafenib (Tafinlar) can be single treatment medications. RET inhibitors including selpercatinib (Retevmo) and pralsetinib (Gayreto), and cabozantinib (Cometriq, Cabometyx) act against RET, ROS1, MET, and VEGF. MET inhibitors including capmatinib (Tabrecta), tepotinib (Tepmetko), and crizotinib (Xalkori) act against MET, ALK, and ROS1. HER2-directed drugs include fam-trastuzumab deruxtecan-nxki (Enhertu) and ado-trastuzumab emtansine (Kadcyla). The examples above include sotorasib (Lumakras) for NSCLC K-Ras G12C mutation,260 which can couple with upstream SHP2 inhibitor vociprotafib (RMC-4630). Notably, sotorasib plus panitumumab (Vectibix) are used in refractory colorectal cancer.261
Experimental approaches explore heterogeneity,262 EMT,263 antibody-drug conjugate in advanced HER2-mutant NSCLC, combinations,264,265 immunotherapy modulation, including immune checkpoint,265 RTK resistance, mutations, adverse effects, signaling,266 metastasis, genome instability, prognosis, apoptosis, degradation, and more, providing an overview. Recently, a combined inhibition of K-RasG12C and mTORC1 kinase was also offered.266
Additional solid tumors that can be aggressive include HCC or liver cancer, colorectal cancer, breast cancer, and prostate cancer.
Leukemia
Acute leukemia (AL) can be an example of a blood cancer267 and includes acute lymphoblastic leukemia (ALL) and acute myeloid leukemia (AML). Unlike chronic lymphocytic leukemia (CLL) and chronic myelogenous leukemia (CML), which progress slowly, both ALL and AML progress rapidly. AL results from clonal proliferation of myeloid and lymphoid progenitor cells.268 Acute promyelocytic leukemia (APL, the most curable AML subtype) has been labeled a paradigm for targeted differentiation therapy.269 APL results from a t(15;17)(q22;q21) translocation that fuses RARA (a gene encoding retinoic acid receptor α, RARα) on 17q21 to PML (a gene encoding promyelocytic leukemia protein, PML) on 15q22. It generates two fusion genes,270 which express PML–RARα and a reciprocal RARα–PML, an aberrant retinoid receptor. PML–RARα heterodimerizes with the retinoid X receptor (RXR), binding to retinoic acid-responsive elements, stalling myeloid differentiation. The resulting high population of promyelocytes expresses and binds factor VII (proconvertin), activating factor X (Stuart-Prower factor) and factor IX (antihemophilic factor B), and pro-coagulant states.271,272
AML is the most aggressive adult leukemia, with clonal differentiation arrest of progenitor or precursor hematopoietic cells.273 It initiates in the myeloid cells of the bone marrow and spreads into the blood. Overexpression of homeobox protein Hox-A9 (HOXA9), a pioneer transcription factor in myeloid and B progenitor cells, is key to its aggressiveness.274 Under physiological conditions, the activity of HOXA9 is primarily regulated by histone methyltransferase MLL3/MLL4. HOXA9-binding sites are enriched with H3K4me1 and H3K27ac, with low levels of H3K27me3. Increased H3K27me3 and H3K4me3 depress binding.275 Deletion of MLL3/MLL4 methylases prevents H3K4 methylation and HOXA9-promoted leukemia.274 The primary H3K27me3 “writer” protein, enhancer of zeste homolog 2 (EZH2), is a component of the polycomb repressor complex 2 (PRC2). Mutations in epigenetic factors are frequent in hematological malignancies.276,277 Lower expression of EZH2 and chromosomal translocations leading to MLL-fusion proteins activate HOXA9 expression through dysregulated chromatin modification.278 Fusion with nuclear pore complex protein NUP98 and overexpression of CDX2 and CDX4 further upregulate HOXA9.276 Thus, AML’s overexpressed transcription factors, and its altered epigenetic landscape and associated transcriptomic program not only harness an embryonic transcriptional development program. It also clarifies the distinction between the milder APL and the aggressive AML.
An excellent overview of targeting mutations in cancer includes a useful detailed compilation of genetic indications for targeted therapy in cancer and the timeline of their approval by the FDA.255 We refer the readers to this review and to the websites listed here for more detailed information. On a different note, it was observed that common anti-cancer therapies induce somatic mutations in stem cells of healthy tissue.279 As the authors note, being toxic to cells, chemotherapies can also increase the mutational burden of long-lived normal stem cells increasing the risk for developing second cancers.
Experimental exploratory approaches include PI3K inhibition,280 signaling,280,281 transcription factors, programmed cell death,282 tyrosine kinase,283 immune modulation,284 prognosis, and more.285
Learning from aggressive cancers
Above, we discussed examples of highly aggressive, often deadly cancers. Our goals were two-fold: First, since they are at the tip of cancer aggressiveness, we deemed that they could inform key properties that are responsible for cancers’ capabilities, enabling forecasting cancers evolution. Second, more importantly, this knowledge may suggest a more informed pharmacological regimen, especially involving treatment combinations. Cancer is a complex disease driven by numerous factors, including environment, lifestyle, and genetic make-up.286,287,288,289 Some cancers are established to be rooted in familial predispositions, such as breast290,291,292,293,294 and colon295,296 cancers. Others are not. Yet, this premise may change with future discoveries. Cancer is not a single-mutation event.26,297,298,299,300,301,302 Multiple mutations are involved, with the minimal number still being debated. Rare, unidentified familial mutations can be expected to preexist, and to collaborate with emerging mutations during life, contributing to cancers, which to date are not thought to be familial. Such scenarios were also proposed to be the case in neurodevelopmental disorders and pathologies.302,303,304,305,306,307,308
Here we ask what we can learn from highly aggressive cancers. Especially, can they inform the foundational hallmark—on the molecular level—underlying them? In 2000, Hanahan and Weinberg articulated the hallmarks of cancer.309 A decade later they expanded the list.310 The impact on cancer biology was massive, as can be assessed by the tens of thousands works that related to them (e.g., see refs. 310,311,312,313,314,315,316). Hanahan and Weinberg created a monumental, visionary and practical, framework that organizes the distinctive properties evolved by cancer cells, thereby conceptualizing cancer biology.309 The six hallmarks of cancer that they laid out in 2020 included evading apoptosis, self-sufficiency in growth signals, insensitivity to antigrowth signals, sustained angiogenesis, tissue invasion and metastasis, and limitless replicative potential. These capture perpetuating proliferative signaling through enhanced activity of the signal transduction pathways by mechanisms including mutations, amplifications, evading growth suppressors, resisting cell death, enabling replicative immortality, inducing angiogenesis, and activating invasion and metastasis.309 In 2011, they added cancer abetting genome instability and mutations and tumor-promoting inflammation as enabling characteristics, and deregulating cellular energetics and avoiding immune destruction as emerging hallmarks.310 These capture multiple hallmark functions, alongside reprogramming energy metabolism and evading immune destruction.
The hallmarks underscore the significance of cancer rewiring normal developmental programs, adapting and sabotaging embryonic regulated cell proliferation, migration, polarity, apoptosis, and differentiation,317,318,319,320,321 and deceiving homeostasis.13,322 Cancer cells acquire the potential to proliferate and survive, with rare mutations and epigenetic alterations selected in drug resistance.323,324,325,326,327,328,329 Subsequently, they were further discussed by Hanahan,330 who explained his perspective, that in conceptualizing the hallmarks, they produced a heuristic tool that can clarify the complexity of cancer—its phenotypes and genotypes.331 That is, the set of hallmarks produced the simplest, and most important, facts and theories relating to cancer, which are not obvious, but detectable. He further elaborated on cancer’s phenotypic plasticity, the disrupted capability of differentiation, and the significance of epigenetic reprogramming and the microbiomes, both not included earlier, but established in the literature,332,333,334 as well as senescence and the tumor microenvironment.93,335,336,337,338,339
A modified hallmarks model suggested that invasion and metastasis should be the center of attention,340 making it a potentially highly ranked therapeutic target. The model argued that therapies aiming at other hallmark-specific mechanisms affect cell viability directly, however, the cancer cells population still undergoes selection and Darwinian evolution. On the other hand, obstructing metastasis stalls seeding—thus distant colonies, which is responsible for most cancer mortalities.
In line with this view, discussions of the hallmarks341 highlighted sustained proliferative signaling. It was argued that with cancer cells stimulating their own growth, they no longer depend on external signals—but largely on two major proliferative pathways: PI3K/AKT/mTOR and MAPK/ERK. Both consist of kinase cascades, and under physiological conditions, both are activated by stimulated RTKs. PI3K, AKT, and mTOR kinases are crucial in metabolism, growth, proliferation, survival, transcription, and protein synthesis.342,343,344,345,346,347,348 These kinases are also major regulators of cell survival by blocking apoptosis. The MAPK pathway is the major signaling cascade in cell proliferation.15,349,350,351,352,353 It transmits and augments survival signals from the cell surface to the nucleus. It is also vital to cell growth. Its stimulation activates ERK, p38 MAPK, and JNK, resulting in activation of transcription factors, including c-Jun, Myc, Elk, ETS, and ATF, and in turn, cell growth, survival, repair, and proliferation. Temporal and spatial topography of cell proliferation in cancer observed that proliferative architecture is organized at two spatial scales: large domains, and smaller niches enriched for specific immune lineages,354 capturing clinically significant features of cancer proliferation.351,355,356,357,358,359
Molecular principles underlying aggressive cancers
What is then the foundational principle—on the molecular, conformational level—underlying aggressive cancers? Going back to the aggressive cancers’ examples discussed above, all harness mechanisms unleashing immense amplification of destructive signaling. Their actions are amplified through topographies such as their organizations in super-enhancers, fusions, alteration of copy numbers and making their enhancer and promoters accessible through epigenetics modulations (Fig. 1). Some transcription factors whose contributions are massive are dubbed master factors, others, pioneers. We also see extensive mutations in multiple relevant RTKs, and their amplifications. Mutant, dysregulated RTKs pump down high, incessant signaling volumes. Other key upstream nodes, like Ras, are mutated as well. These commonly co-occur multiple times as described in the examples. Taken together, aggressive cancers are characterized by heavy loads of active proteins, especially kinases, key signaling nodes, and missing, or inactive tumor suppressors. Under physiological conditions, mitogen-promoted signaling strength and duration can control cell cycle decisions; to proliferate or differentiate.12 Strong and short encode proliferation, weaker and sustained differentiation. In cancer, the scenarios above lead to hyper-strong, long-lived proliferation. Signaling has been described phenomenologically, as propagating from one node to another.14,360,361,362,363,364,365 Its impact on metabolism and as therapeutic targets has been reviewed, and signaling-based combinatorial treatments were offered.57,58,59,366 How cancer cells can send wrong signals was also deliberated, examples of signaling in cancer were described,367,368,369,370,371 and an overview of signaling pathways in cancer was provided. This overview highlighted the consequences of overexpressed signaling components causing dysregulation of cellular signaling, through cross talk.372,373 At the fundamental level, we formulated a signaling by-the-numbers model, which also considers the cell type (a skin cell differs from a brain cell) and timing (here, of cancer evolution). Such model considers the total, absolute numbers of active conformations of the mutant protein, the types, and locations of all its co-existing mutations, and the expression levels of specific isoforms of genes and regulators of proteins in the pathway. We suggest that tumors harboring massive, cancer-promoting, catalysis-primed oncogenic payloads of active protein conformations, especially through certain overexpression scenarios, are predisposed aggressive tumor candidates. The stronger the combination of driver mutations, and the bigger the number of the respective proteins, and of those who transmit the signal down the pathway, the stronger the signal transduction.
This leads us to designate superstrong proliferative signaling as the molecular principle underlying aggressive cancers, making it a foundational hallmark. Superstrong signaling boosts invasion and metastasis, making it a major cancer capability, providing the basis on the molecular level of these fundamental cancer phenotypes.340 Tissue invasion and metastasis, and limitless replicative potential capabilities require formidable activity of the signal transduction pathways.
As to pharmacology, it substantiates the call to invest in invasion and metastasis. Glioblastoma is an invasive brain tumor. Other highly aggressive cancers metastasize. Yet, to date few treatments were developed,374 including few suggested combinations.375,376,377,378,379 Often, metastases drugs in the chemotherapy cocktail are the same as those against the corresponding primary tumors, rather than metastases, as in metastatic breast cancer. Drugging proteins associated with metastasis is challenging. One example relates to late-stage metastatic melanoma, recently suggested to be hallmarked by low expression of postsynaptic cell adhesion molecule neuroligin 4X (NLGN4X), a suppressor, leading to HIF1A accumulation, and acquisition of migratory properties.380
Questions and clinical implications of the model
Here we suggested a model of “cancer aggressiveness by-the-numbers.” Questions and implications for the clinics include:
-
(1)
Is there a numerical threshold for high aggressiveness? Determining a threshold is a vastly important aim—but to date, a still unattained challenge. Having such a number could help diagnose emerging cancer mutations and proliferation prior to observable phenotypic change.
-
(2)
Would the threshold be the same for different types of cancers? As to threshold consistency across cancer types, since the number depends on the strength of the mutations, extent of overexpression (e.g., gene duplication, super-enhancers, combination of genetic and epigenetic events), and the identity of the targeted genes, which vary across cancer types and evolve with cancer stages, we expect that thresholds will vary for different cancer types.
-
(3)
What qualifies as an active (oncogenic) conformation? An active protein conformation is the shape that the protein takes when it is biologically active. The process of changing shape is a conformational change.
-
(4)
How to count the active conformations? The answer is two-fold. Both ways provide estimates. The first way is measuring indirectly, by protein activity, similar to activation of proteins by driver mutations. Driver mutations activate oncogenic proteins. They lead to a conformational change from the inactive to the active conformation. Stronger mutations are likely to have lower kinetic barriers in the transitions and more stable conformations. They are expected to have a higher fraction in the active state, making the mutations stronger. Experiments may identify a stronger signaling. The second way is direct measurement, by e.g., mapping the protein conformational ensemble by a spectroscopic method, like NMR spectroscopy, or by long timescale molecular dynamic simulations.
-
(5)
Are all active conformations treated equally when counted? All active conformations are treated equally. Chemistry tells us: for a protein there is a single active state, where all catalytic groups are precisely coordinated (the right angle and the right distance) for the catalytic reaction. There can however be substate variations, reducing catalytic efficiency.
-
(6)
How are super-enhancers quantified? The expression of super-enhancers can be measured by e.g., chromatin immunoprecipitation and sequencing (ChIP-Seq).
-
(7)
How is protein overexpression assessed? Protein overexpression can be assessed using a variety of methods, for example, Western blotting, In-cell western assays, and comparing normal levels with those in the mutant cell.
-
(8)
Are tumor suppressors included in this count, and if so, how? No, they are not included. Repressors work by tamping down on expression, thus active conformations.
-
(9)
Is there verification and examples for the model? Verification comes from analysis of known mechanisms of aggressive cancers. This indicates that they acquire activating mutations, overexpression, gene fusion, gene duplication, epigenetics, and more. All lead to proliferation, the outcome of strong signaling through a higher number of active proteins, e.g., in the MAPK and PI3K/AKT/mTOR pathways. Inhibitors depress catalytic activity. As to examples, one protein that we worked on recently is Bcr-Abl in chronic myeloid leukemia (CML).381,382 C-Abl is a protein kinase in the cytoplasm. The kinase domain is in the C-terminal region. It is allosterically regulated by myristoyl (a lipidic post-translational modification), which is covalently linked at its N-terminal tail. Bcr is a transmembrane protein on the surface of B cells. It is attached to the membrane through its N-terminal domains. Under normal conditions, Abl is autoinhibited by the myristoyl which docks into its pockets at the C-lobe kinase domain, retaining Abl in an inactive conformation. In leukemia, Abl’s C-terminal domain is fused with Bcr’s N-terminal domain, removing its capacity to be autoinhibited and making it membrane associated, in a constitutively active state. Another type of example relates to the relative concentrations of the active conformations and clinical diagnosis and treatment.302 In this example, SHP2 clinical phenotype—cancer or RASopathies—can be predicted by mutant conformational propensities. SHP2 phosphatase promotes full activation of the RTK-dependent Ras/MAPK pathway. Its mutations can drive cancer and RASopathies, a group of neurodevelopmental disorders. We asked how same residue mutations in SHP2 can lead to both cancer and RASopathies phenotypes, and whether we can predict what the clinical outcome will be. We observed that SHP2 clinical phenotype can be predicted by mutant conformational propensities. High propensity of the active conformation was associated with cancer, lower propensity with RASopathy, offering structural guidelines for identifying and correlating mutations with clinical outcomes, and a drug strategy.
-
(10)
What are the limitations of this proposed model? Changes in the number of active conformations can also take place under physiological conditions. Among these, cell type-specific expression of the protein, and of other proteins in the respective pathway, timing of activation (during embryonic development or sporadic emergence), are pivotal. Further, activation is the result of allosteric events. There can be multiple such events unrelated and related to cancer, for example, effects of the cellular environment, metabolites, and the dependence on the combination of driver mutations, in the respective proteins, and in those which transmit the signal down the pathway.
-
(11)
How this model can be used in clinical practice? Usefulness in clinical practice is the goal. An excessively high number of active conformations lead to more active proteins. Higher activity is associated with oncogenic events in cancer evolution. Clinically, this suggests that signaling measurements can be useful in early detection of cancer in tissue samples (biopsies), in addition to visual microscopic examination, which is not quantitative.
Clinical targeting aggressive cancers
Drugs and drug combinations targeting aggressive cancers
There has been clinical progress in targeting the molecular mechanisms and therapeutic targets of aggressive cancers, including clinical trials resulting in FDA-approved drugs. In addition to neuroblastoma, glioblastoma, PDAC, NSCLC, and leukemia, which we discussed above, additional aggressive cancer types, such as breast cancer, liver cancer and CRC have also been actively challenged and are described in this section. To cite examples, in early-stage triple-negative breast cancer, immunotherapy before surgery was determined as able to improve the prognosis, becoming standard-of-care. For extensive NSCLCs, the FDA approved a combination of immunotherapy and chemotherapy. These and other clinical advances have been included in the report of the American Society of Clinical Oncology.
FDA-approved drugs for neuroblastoma383 include cyclophosphamide, dinutuximab, doxorubicin hydrochloride, eflornithine hydrochloride (or DFMO, difluoromethylornithine), naxitamab-gqgk, and vincristine sulfate (Table 1). Dinutuximab is the first therapy specifically approved for pediatric high-risk neuroblastoma in combination with interleukin-2 (IL-2 or aldesleukin), granulocyte-macrophage colony-stimulating factor (GM-CSF), and isotretinoin.384,385 Naxitamab-gqgk was approved for the treatment of neuroblastoma with high refractory or relapse risk in bone marrow and/or bone in combination with GM-CSF.386 eflornithine was approved for reducing the risk of relapse in pediatric patients with high-risk neuroblastoma who are in remission and have completed multi-agent, multi-modality therapy.387,388 Drug combinations used in neuroblastoma include BuMel (busulfan + melphalan hydrochloride) and CEM (carboplatin + etoposide phosphate + melphalan hydrochloride).383
For glioblastoma patients, FDA-approved treatments of high-grade malignant gliomas389 include bevacizumab, carmustine, lomustine, temozolomide, and vorasidenib (Table 2). A form of carmustine contained in a wafer, carmustine implant (Gliadel), was also approved for the treatment of glioblastoma multiforme and is applied directly during surgery.390 Vorasidenib was recently approved, the first approval in decades, for patients with grade 2 gliomas harboring IDH1 or IDH2 mutations.391 Drug combinations used in certain types of brain tumors include PCV (procarbazine hydrochloride + lomustine + vincristine sulfate).389
The list of drugs approved for pancreatic cancer patients225 includes multiple drugs (Table 3). These drugs approved for pancreatic cancer are often also administered against other cancers, such as breast and lung cancer. In addition to pancreatic cancer, paclitaxel, everolimus, capecitabine, 5-fluorouracil, gemcitabine hydrochloride, and olaparib are also approved for breast cancer. Similarly, paclitaxel, everolimus, erlotinib hydrochloride, and gemcitabine hydrochloride are also approved for NSCLC. Drugs approved for the treatment of tumors in the digestive system include everolimus, capecitabine, 5-fluorouracil, mitomycin, and sunitinib malate. Drugs approved for the treatment of other cancers include everolimus and sunitinib malate for kidney cancer; everolimus for brain cancer; capecitabine and 5-fluorouracil for CRC; gemcitabine hydrochloride and olaparib for ovarian cancer; olaparib for prostate cancer; and mitomycin for urothelial cancer. Drug combinations used in pancreatic cancer include FOLFIRINOX (leucovorin calcium + fluorouracil + irinotecan hydrochloride + oxaliplatin), GEMCITABINE-CISPLATIN (gemcitabine hydrochloride + cisplatin), GEMCITABINE-OXALIPLATIN (gemcitabine hydrochloride + oxaliplatin), and OFF (oxaliplatin + fluorouracil + leucovorin calcium (folinic acid)).225
For NSCLC, the FDA has approved a large number of small molecule and monoclonal antibody drugs (Table 4).392 These small molecule drugs target K-RasG12C (adagrasib, sotorasib); the kinase domain of multiple RTKs, including the EGFR family, MET, ALK, ROS1, IGF1R, FLT3, RET, the Trk family and the EML4-ALK fusion protein (afatinib dimaleate, alectinib, brigatinib, capmatinib hydrochloride, ceritinib, crizotinib, dacomitinib, entrectinib, erlotinib hydrochloride, gefitinib, lazertinib mesylate hydrate, lorlatinib, osimertinib mesylate, pralsetinib, repotrectinib, selpercatinib, tepotinib hydrochloride); B-Raf mutants (dabrafenib mesylate, encorafenib); MEK1/2 (binimetinib, trametinib dimethyl sulfoxide); mTOR (everolimus); tubulin (docetaxel, paclitaxel albumin, vinorelbine tartrate); enzymes responsible for nucleotide synthesis (methotrexate sodium, pemetrexed disodium); and DNA (doxorubicin hydrochloride, gemcitabine hydrochloride). The monoclonal antibodies target the extracellular domain of RTKs on the surface of tumor cells, including EGFR, HER2 and VEGFR (amivantamab-vmjw, bevacizumab, necitumumab, ramucirumab, trastuzumab) and block the receptors, programmed cell death 1 (PD-1) and cytotoxic T-lymphocyte antigen 4 (CTLA-4) on the surface of T cells (cemiplimab-rwlc, ipilimumab, nivolumab, pembrolizumab, tremelimumab-actl) and the ligand PD-L1 on the surface of tumor cells (atezolizumab, durvalumab). Drug combinations used in NSCLC include CARBOPLATIN-TAXOL (carboplatin + paclitaxel) and GEMCITABINE-CISPLATIN (gemcitabine hydrochloride + cisplatin).392
The FDA has approved a significant number of drugs for the treatment of leukemia, including ALL, AML, CLL and CML.393 Furthermore, a number of drugs have been approved for use in the treatment of AML only (Table 5). The FDA also approved a combination form of daunorubicin hydrochloride and cytarabine contained inside liposome (Vyxeos) for AML. Idarubicin is a structural analog of daunorubicin with the same mechanism of action. Drug combinations used in AML include ADE (cytarabine + daunorubicin hydrochloride + etoposide phosphate).393
In the case of breast cancer, scores of drugs have been granted approval,394 with the list of these drugs too extensive to enumerate. Some of the drugs approved to treat breast cancer are also used to treat other types of cancer, including cyclophosphamide and doxorubicin hydrochloride for neuroblastoma; 5-fluorouracil, capecitabine, everolimus, gemcitabine hydrochloride, olaparib and paclitaxel albumin for pancreatic cancer; docetaxel, doxorubicin hydrochloride, everolimus, gemcitabine hydrochloride, methotrexate sodium, paclitaxel albumin, pembrolizumab and trastuzumab for NSCLC. The individual drugs in the combinations are FDA-approved. However, the drug combinations are commonly not approved, although widely used. The most recent FDA approved drug is ribociclib succinate (Kisqali) with an aromatase inhibitor for the adjuvant treatment of adults with hormone receptor (HR)-positive, HER2-negative stage II and III early breast cancer. Drug combinations used in breast cancer includes AC (doxorubicin hydrochloride + cyclophosphamide), AC-T (doxorubicin hydrochloride + cyclophosphamide + paclitaxel), CAF (cyclophosphamide + doxorubicin hydrochloride + fluorouracil), CMF (cyclophosphamide + methotrexate + fluorouracil), FEC (fluorouracil + epirubicin hydrochloride + cyclophosphamide), and TAC (docetaxel + doxorubicin hydrochloride + cyclophosphamide).394
Liver cancer is not included in our detailed discussion of aggressive cancers above, but it can be highly aggressive. Liver cancer is often unamenable to surgery since by the time it is diagnosed it has already spread or is intertwined with blood vessels. For liver cancer, the FDA approved atezolizumab (Tecentriq) and bevacizumab (Avastin).395 As expected, clinical studies indicated that their combination is better than the single therapies. Atezolizumab is an immune checkpoint inhibitor. Bevacizumab prevents growth of new blood vessels. The updated list of drugs approved for liver cancer includes the atezolizumab and bevacizumab discussed above.396 Drugs approved to treat HCC, a type of liver cancer, are also used to treat other types of cancer, including bevacizumab for glioblastoma; atezolizumab, bevacizumab, durvalumab, ipilimumab, pembrolizumab, ramucirumab and tremelimumab-actl for NSCLC. The drugs ivosidenib and pemigatinib, which are approved to treat AML, are also approved to treat cholangiocarcinoma (CCA), a rare and aggressive cancer that forms in the bile ducts. Developing from cells in the liver, liver cancer can be especially aggressive. There are a few types determined by the liver cell where it originated.397 The major types include (1) HCC, which is the most common. Eighty % of the liver cells are hepatocytes. Severe liver damage, or cirrhosis is believed to be the origin for this cancer. (2) Cholangiocarcinoma is a bile duct cancer. (3) Fibrolamellar carcinoma is very rare, occurring in people with healthy livers. (4) Hepatoblastoma is an extremely rare type of liver cancer but common in young children. (5) Liver angiosarcoma also rare, forms in blood cells or lymph vessels, part of your immune system.
CRC develops slowly, and if caught early, is treatable. However, it is the second deadliest cancer in the U. S., with its metastasis posing a challenge, primarily due to limited therapeutic options. Optimal treatment depends on the tumor type. High-grade large cell and small cell neuroendocrine tumors are aggressive. Metastatic CRC (mCRC) tends to be relatively more lethal than primary CRC. 30% of colorectal cancers are mesenchymal mCRC, with hyaluronan accretion a key step in mCRC tumors formation, especially under low levels of PKCζ and PKCι, resulting in worse prognoses.398 A combination of anti-PD-L1 and anti-CTLA-4 antibodies with hyaluronidase, which breaks down hyaluronan, has been investigated as a potential treatment for targeting CRCs that metastasized to the liver, which is a common occurrence, making hyaluronidase an important agent for immunotherapy success. A recent review398 discussed the genetic alterations associated with advanced CRC and metastasis as well as the role of cellular heterogeneity and plasticity in drug resistance and metastasis and how the tumor microenvironment impacts it. However, beyond generic concepts, much is still not understood, hampering effective pharmacologic treatment. The FDA has approved a number of drugs for the treatment of colon and rectal cancer.399 Broadly, mesenchymal stromal cells (MSCs) are constituents of tumor stroma.400 They promote a favorable environment for tumor progression, proliferation, and metastasis, clarifying why metastatic mCRC is often lethal.
Above, we overviewed currently available and prescribed single drugs, and drug combinations. In the interest of brevity, we did not detail the ages of the patients, prior treatments, etc., which are provided at the links that we cited.
Clinical research progress: treatment design
Here we take up the challenging question of directions for future research and treatment designs for highly aggressive cancers. We consider the rational, available relevant tools and data, and expanding their applications. The question is which combination of targets to consider, which can guide drug combinations. With this goal in mind, we suggest considering that primary cancers are tissue specific, and that highly aggressive cancers present an extensive heterogeneity, proliferation, and metastases. This argues for harnessing spatial single-cell transcriptomic to identify microclones, which are subpopulations of cancer cells that exist within tumors, and analysis of the cancer-specific metastases to identify the tissues that the cancer favors for its metastases to settle in.401,402,403 Data inform that specific cancers prefer certain host tissues. This leads us to suggest that besides comprehensive searches to identify the metastases, analyses of the overexpressed genes be carried out, as these can provide clues as well. Early data by Axelsen et al.404 observed that at least some cancers tend to metastasize in tissues which express those overexpressed genes. They suggested, and we believe, that this is likely to be a general trend. These data can couple with statistical trends of cancer-specific co-occurring driver mutations, mapped on their harboring proteins and pathways. Drug resistance commonly exploits parallel pathways and bypasses engaged through pathway crosslinkages, which are overflooded in aggressive cancers. Spatial single-cell transcriptomic is crucial to diagnose overexpression, which commonly involves epigenetic alteration, not identifiable via mutational analyses.
In addition to identifying pathways by mapping of mutations, pathway-based approaches are currently being used to study cancer in biological networks, including Gene Set Enrichment (GSE), which identifies pathways enriched for genes with altered expression;405 Over-Representation Analysis (ORA), which identifies pathways with more differentially expressed genes than would be expected by chance; Function Class Scoring (FCS), which identifies disease pathways based on the aggregate of their gene expression values; Differential Causal Effects (DCE), which identifies dysregulated signaling pathways in cancer cells and compares normal and cancer cells using the statistical framework of causality;406 and CTpathway, which identifies cancer target pathways in early-stage tissues and blood samples.407
Collectively, our treatment designs against aggressive cancers innovates by offering research involving analyses of cancer-specific metastases, and primary and metastases-specific tumor single-cell spatial data, and further supporting broad pathway analyses along with transcription factors and epigenetic modulators. Our on-going work harnesses an innovative pathway-mapping strategy.
Conclusions
Here we asked: What are the underlying molecular principles of aggressive cancers? What distinguishes them from the less aggressive, more treatable ones? And crucially, can their molecular hallmarks offer clues to anti-cancer drug combination strategies targeting their ultra-strong drug resistance?
Our review took up these questions, noting that whereas several factors can be in play,1,2,3,4,5 the accelerated growth and malignancy of aggressive cancers point to a massive payload of primed, catalysis-ready states resulting from unmitigated overexpression of oncogenic proteins, and high load of activation mutations. High population of these oncogenic signaling nodes activates not only its functionally intended pathways, but multiple oncogenic pathways, as shown in the case of PIK3CA,18 fostering heterogeneity and proliferation. In glioblastoma, the mutations alter EGFR dimer formation, attenuating ligand bias, corrupting downstream signaling.408 Mutations that alter the transcriptional landscapes, perturb the protein interactome,409,410 increase cancer heterogeneity and aggressiveness. High oncogenic mutational loads, and overexpression, break homeostatic physiological feedback mechanisms.322,411
The National Cancer Institute defined aggressive cancer as one that spreads rapidly, despite severe treatment. Here we offer the underlying principle on the fundamental structural level: Aggressive cancers are those harboring an exceedingly large population of proteins critical in signal transduction in major proliferation pathways in their active, catalysis-prone states. That is, the structures of these dominant signaling proteins are in their active conformations, primed to execute their signaling roles.412 This definition implicitly articulates overexpression of master transcription factors, transcription factors with gene fusions, copy number alterations, and dysregulation of the epigenetic codes, especially in super-enhancers. High mutation loads of vital upstream signaling regulators (Fig. 1), such as RTKs and K-Ras, also fits into this definition. A combination of overexpression of multiple such proteins, epigenetic alterations, and activating mutations of major upstream signaling nodes can be decisive, suggesting that acquisition of these attributes can serve as hallmark of aggressive cancers. Their impact can be quantified by signaling by-the-numbers scenarios, where the numbers refer to the number of active state conformations.13
While the same protein may populate many cancers, its abundance and the predominant mutation may vary, depending on the cell type, state, cell history, the background and evolving mutational load, with epigenetics and chromatin accessibility of the respective genes playing major roles.43,413 Heterogeneity implies that resistance mutations are likely to exist prior to treatment. As discussed here, drugs harnessed to target aggressive cancers may overlap, but also be distinct aiming at the specific proteins and their more frequent mutations.
Forecasting these can assist the attending oncologist.24,414,415 As to anti-cancer drug combination strategies, mapping the drugs onto the respective signaling pathways, may indicate their type of complementarity, and how it might be improved.57,58 Coupling it with pathway mapping of the oncogenic proteins whose expression indicates substantial deviation from that of normal tissues, and those that are highly mutated, may better inform combinatorial selection. These should be coupled with drugs targeting epigenetics regulators, master transcription factors, fused transcription factors, and those with copy number variations.
To conclude, here we propose that the absolute number of active (oncogenic) conformations that the cancer harbors are a foundational hallmark of its aggressiveness (Fig. 1). The higher the number—the more overspilled the signaling—the higher the heterogeneity. In aggressive cancers the number is extremely high. We call this hallmark “cancer aggressiveness by-the-numbers”. We designate it “foundational” since it is expressed on the molecular level by conformational distributions, the most fundamental physical-chemical attribute of biomacromolecules. Dynamic conformational, or ensemble, propensities decide molecular, and cell function.37,38
References
Martinez-Useros, J., Martin-Galan, M., Florez-Cespedes, M. & Garcia-Foncillas, J. Epigenetics of most aggressive solid tumors: pathways, targets and treatments. Cancers 13, 3209 (2021).
Saggese, P. et al. Glucose deprivation promotes pseudohypoxia and dedifferentiation in lung adenocarcinoma. Cancer Res. 84, 305–327 (2024).
Watson, A. W. et al. Breast tumor stiffness instructs bone metastasis via maintenance of mechanical conditioning. Cell Rep. 35, 109293 (2021).
Takahashi, H. et al. Biologically aggressive phenotype and anti-cancer immunity counterbalance in breast cancer with high mutation rate. Sci. Rep. 10, 1852 (2020).
Docrates Cancer Center. The most aggressive cancers: pancreatic, lung and liver cancer show the fastest progression. https://www.docrates.com/en/the-most-aggressive-cancers/ (2023).
Chen, X. et al. Cyclin E overexpression sensitizes triple-negative breast cancer to Wee1 kinase inhibition. Clin. Cancer Res. 24, 6594–6610 (2018).
Xu, T. et al. Overexpression of protein regulator of cytokinesis 1 facilitates tumor growth and indicates unfavorable prognosis of patients with colon cancer. Cancer Cell Int. 20, 528 (2020).
Zhang, R. et al. CXCL5 overexpression predicts a poor prognosis in pancreatic ductal adenocarcinoma and is correlated with immune cell infiltration. J. Cancer 11, 2371–2381 (2020).
He, Y. et al. FOXA1 overexpression suppresses interferon signaling and immune response in cancer. J. Clin. Invest. 131, e147025 (2021).
Wang, Y. F. et al. Overexpression of Stat3 increases circulating cfDNA in breast cancer. Breast Cancer Res. Treat. 187, 69–80 (2021).
Bodea, I. C. et al. HER2 overexpression in periampullary tumors according to anatomical and histological classification—a systematic review. J. Pers. Med. 14, 463 (2024).
Nussinov, R., Zhang, W., Liu, Y. & Jang, H. Mitogen signaling strength and duration can control cell cycle decisions. Sci. Adv. 10, eadm9211 (2024).
Nussinov, R., Tsai, C. J. & Jang, H. A new view of activating mutations in cancer. Cancer Res. 82, 4114–4123 (2022).
Nussinov, R., Tsai, C. J. & Jang, H. Allostery, and how to define and measure signal transduction. Biophys. Chem. 283, 106766 (2022).
Wang, P. et al. Erianin suppresses constitutive activation of MAPK signaling pathway by inhibition of CRAF and MEK1/2. Signal Transduct. Target Ther. 8, 96 (2023).
Xiao, J. et al. Integrating spatial and single-cell transcriptomics reveals tumor heterogeneity and intercellular networks in colorectal cancer. Cell Death Dis. 15, 326 (2024).
Zuo, C. et al. Elucidating tumor heterogeneity from spatially resolved transcriptomics data by multi-view graph collaborative learning. Nat. Commun. 13, 5962 (2022).
Gustin, J. P. et al. Knockin of mutant PIK3CA activates multiple oncogenic pathways. Proc. Natl Acad. Sci. USA 106, 2835–2840 (2009).
Liu, P., Cheng, H., Roberts, T. M. & Zhao, J. J. Targeting the phosphoinositide 3-kinase pathway in cancer. Nat. Rev. Drug Discov. 8, 627–644 (2009).
Yuan, W. et al. Conditional activation of Pik3ca(H1047R) in a knock-in mouse model promotes mammary tumorigenesis and emergence of mutations. Oncogene 32, 318–326 (2013).
Wu, X. et al. Activation of diverse signalling pathways by oncogenic PIK3CA mutations. Nat. Commun. 5, 4961 (2014).
Robertson, S. C., Tynan, J. A. & Donoghue, D. J. RTK mutations and human syndromeswhen good receptors turn bad. Trends Genet. 16, 265–271 (2000).
Du, Z. & Lovly, C. M. Mechanisms of receptor tyrosine kinase activation in cancer. Mol. Cancer 17, 58 (2018).
Nussinov, R., Jang, H., Nir, G., Tsai, C. J. & Cheng, F. A new precision medicine initiative at the dawn of exascale computing. Signal Transduct. Target Ther. 6, 3 (2021).
Nussinov, R., Tsai, C. J. & Jang, H. How can same-gene mutations promote both cancer and developmental disorders? Sci. Adv. 8, eabm2059 (2022).
Nussinov, R., Yavuz, B. R. & Jang, H. Single cell spatial biology over developmental time can decipher pediatric brain pathologies. Neurobiol. Dis. 199, 106597 (2024).
Grudzien, P., Jang, H., Leschinsky, N., Nussinov, R. & Gaponenko, V. Conformational dynamics allows sampling of an “Active-like” state by oncogenic K-Ras-GDP. J. Mol. Biol. 434, 167695 (2022).
Xie, T., Saleh, T., Rossi, P. & Kalodimos, C. G. Conformational states dynamically populated by a kinase determine its function. Science 370, eabc2754 (2020).
Tong, M. & Seeliger, M. A. Targeting conformational plasticity of protein kinases. ACS Chem. Biol. 10, 190–200 (2015).
Parker, M. I., Meyer, J. E., Golemis, E. A. & Dunbrack, R. L. Delineating the RAS conformational landscape. Cancer Res. 82, 2485–2498 (2022).
Möbitz, H. Conformational bias: a key concept for protein kinase inhibition. https://www.europeanpharmaceuticalreview.com/article/11289/conformational-bias-a-key-concept-for-protein-kinase-inhibition/ (2012).
Petrovic, D., Risso, V. A., Kamerlin, S. C. L. & Sanchez-Ruiz, J. M. Conformational dynamics and enzyme evolution. J. R. Soc. Interface 15, 20180330 (2018).
Lo, C. H., Huber, E. C. & Sachs, J. N. Conformational states of TNFR1 as a molecular switch for receptor function. Protein Sci. 29, 1401–1415 (2020).
Tang, W. S., Zhong, E. D., Hanson, S. M., Thiede, E. H. & Cossio, P. Conformational heterogeneity and probability distributions from single-particle cryo-electron microscopy. Curr. Opin. Struct. Biol. 81, 102626 (2023).
Monteiro da Silva, G., Cui, J. Y., Dalgarno, D. C., Lisi, G. P. & Rubenstein, B. M. High-throughput prediction of protein conformational distributions with subsampled AlphaFold2. Nat. Commun. 15, 2464 (2024).
Kawamukai, H. et al. Conformational distribution of a multidomain protein measured by single-pair small-angle X-ray scattering. J. Phys. Chem. Lett. 15, 744–750 (2024).
Nussinov, R., Liu, Y., Zhang, W. & Jang, H. Cell phenotypes can be predicted from propensities of protein conformations. Curr. Opin. Struct. Biol. 83, 102722 (2023).
Nussinov, R., Liu, Y., Zhang, W. & Jang, H. Protein conformational ensembles in function: roles and mechanisms. RSC Chem. Biol. 4, 850–864 (2023).
Vijg, J. & Dong, X. Pathogenic mechanisms of somatic mutation and genome mosaicism in aging. Cell 182, 12–23 (2020).
Risques, R. A. & Kennedy, S. R. Aging and the rise of somatic cancer-associated mutations in normal tissues. PLoS Genet. 14, e1007108 (2018).
Laconi, E., Marongiu, F. & DeGregori, J. Cancer as a disease of old age: changing mutational and microenvironmental landscapes. Br. J. Cancer 122, 943–952 (2020).
Vijg, J. et al. Mitigating age-related somatic mutation burden. Trends Mol. Med. 29, 530–540 (2023).
Nussinov, R., Tsai, C. J. & Jang, H. Anticancer drug resistance: an update and perspective. Drug Resist. Updat. 59, 100796 (2021).
Molteni, E. et al. Prevalence and clinical expression of germ line predisposition to myeloid neoplasms in adults with marrow hypocellularity. Blood 142, 643–657 (2023).
Patel, N. & Calvo, K. R. How I diagnose myeloid neoplasms with germline predisposition. Am. J. Clin. Pathol. 160, 352–364 (2023).
Farinea, G. et al. The role of germline mutations in thoracic malignancies: between myth and reality. J. Thorac. Oncol. 18, 1146–1164 (2023).
Venizelos, A. et al. Germline pathogenic variants in patients with high-grade gastroenteropancreatic neuroendocrine neoplasms. Endocr. Relat. Cancer 30, e230057 (2023).
Demko, N. & Geyer, J. T. Updates on germline predisposition in pediatric hematologic malignancies: what is the role of flow cytometry? Cytometry B Clin. Cytom. https://doi.org/10.1002/cyto.b.22192 (2024).
Orzan, O. A. et al. Cutaneous melanoma and glioblastoma multiforme association-case presentation and literature review. Diagnostics 13, 1046 (2023).
Dantonio, P. M. et al. Exploring major signaling cascades in melanomagenesis: a rationale route for targetted skin cancer therapy. Biosci. Rep. 38, BSR20180511 (2018).
Scarbrough, P. M., Akushevich, I., Wrensch, M. & Il’yasova, D. Exploring the association between melanoma and glioma risks. Ann. Epidemiol. 24, 469–474 (2014).
Barranco, C. Strength lies in difference. Nat. Rev. Cancer 22, 547 (2022).
Bigler, A. Study hints at why an aggressive form of brain cancer is hard to treat. https://www.broadinstitute.org/news/study-hints-why-aggressive-form-brain-cancer-hard-treat (2019).
Neftel, C. et al. An integrative model of cellular states, plasticity, and genetics for glioblastoma. Cell 178, 835–849.e821 (2019).
Simeonov, K. P. et al. Single-cell lineage tracing of metastatic cancer reveals selection of hybrid EMT states. Cancer Cell 39, 1150–1162.e1159 (2021).
Modi, V. & Dunbrack, R. L. Jr. Defining a new nomenclature for the structures of active and inactive kinases. Proc. Natl Acad. Sci. USA 116, 6818–6827 (2019).
Nussinov, R., Yavuz, B. R. & Jang, H. Anticancer drugs: how to select small molecule combinations? Trends Pharm. Sci. 45, 503–519 (2024).
Nussinov, R. & Jang, H. The value of protein allostery in rational anticancer drug design: an update. Expert Opin. Drug Deliv. https://doi.org/10.1080/17460441.17462024.12384467 (2024).
Nussinov, R. et al. Directions to overcome therapy resistance in cancer. Trends Pharm. Sci. 45, 467–471 (2024).
Jaaks, P. et al. Effective drug combinations in breast, colon and pancreatic cancer cells. Nature 603, 166–173 (2022).
Hu, C. et al. Optimizing drug combination and mechanism analysis based on risk pathway crosstalk in pan cancer. Sci. Data 11, 74 (2024).
Gilad, Y., Gellerman, G., Lonard, D. M. & O’Malley, B. W. Drug combination in cancer treatment—from cocktails to conjugated combinations. Cancers 13, 669 (2021).
Mulero-Sanchez, A. et al. Rational combination of SHP2 and mTOR inhibition for the treatment of hepatocellular carcinoma. Mol. Oncol. 17, 964–980 (2023).
Banerjee, D. et al. Lineage specific transcription factor waves reprogram neuroblastoma from self-renewal to differentiation. Nat. Commun. 15, 3432 (2024).
Ge, R., Wang, Z. & Cheng, L. Tumor microenvironment heterogeneity an important mediator of prostate cancer progression and therapeutic resistance. NPJ Precis. Oncol. 6, 31 (2022).
Halliday, A. Tumor heterogeneity: navigating the next frontier in cancer research. https://www.technologynetworks.com/cancer-research/articles/tumor-heterogeneity-navigating-the-next-frontier-in-cancer-research-383052 (2024).
Bhat, G. R. et al. Cancer cell plasticity: from cellular, molecular, and genetic mechanisms to tumor heterogeneity and drug resistance. Cancer Metastasis Rev. 43, 197–228 (2024).
Zhang, A., Miao, K., Sun, H. & Deng, C. X. Tumor heterogeneity reshapes the tumor microenvironment to influence drug resistance. Int. J. Biol. Sci. 18, 3019–3033 (2022).
Caswell, D. R. & Swanton, C. The role of tumour heterogeneity and clonal cooperativity in metastasis, immune evasion and clinical outcome. BMC Med. 15, 133 (2017).
El-Deiry, W. S., Taylor, B. & Neal, J. W. Tumor evolution, heterogeneity, and therapy for our patients with advanced cancer: how far have we come? Am. Soc. Clin. Oncol. Educ. Book 37, e8–e15 (2017).
Anderson, A. C. et al. Spatial transcriptomics. Cancer Cell 40, 895–900 (2022).
Yamada, Y., Haga, H. & Yamada, Y. Concise review: dedifferentiation meets cancer development: proof of concept for epigenetic cancer. Stem Cells Transl. Med. 3, 1182–1187 (2014).
Baraban, E. & Cooper, K. Dedifferentiated and undifferentiated neoplasms: a conceptual approach. Semin. Diagn. Pathol. 38, 119–126 (2021).
National Cancer Institute. Tumor grade. https://www.cancer.gov/about-cancer/diagnosis-staging/diagnosis/tumor-grade (2022).
Martinez-Ruiz, C. et al. Genomic-transcriptomic evolution in lung cancer and metastasis. Nature 616, 543–552 (2023).
Seferbekova, Z., Lomakin, A., Yates, L. R. & Gerstung, M. Spatial biology of cancer evolution. Nat. Rev. Genet. 24, 295–313 (2023).
Park, J. K. et al. Single-cell transcriptome analysis reveals subtype-specific clonal evolution and microenvironmental changes in liver metastasis of pancreatic adenocarcinoma and their clinical implications. Mol. Cancer 23, 87 (2024).
Chaligne, R., Nam, A. S. & Landau, D. A. TARGET-seq takes aim at cancer evolution through multi-omics single-cell genotyping and transcriptomics. Mol. Cell 73, 1092–1094 (2019).
Hoogstrate, Y. et al. Transcriptome analysis reveals tumor microenvironment changes in glioblastoma. Cancer Cell 41, 678–692.e677 (2023).
Li, C. & Wang, J. Quantifying Waddington landscapes and paths of non-adiabatic cell fate decisions for differentiation, reprogramming and transdifferentiation. J. R. Soc. Interface 10, 20130787 (2013).
Wang, J., Zhang, K., Xu, L. & Wang, E. Quantifying the Waddington landscape and biological paths for development and differentiation. Proc. Natl Acad. Sci. USA 108, 8257–8262 (2011).
Iovino, N. & Cavalli, G. Rolling ES cells down the Waddington landscape with Oct4 and Sox2. Cell 145, 815–817 (2011).
Noble, D. Conrad Waddington and the origin of epigenetics. J. Exp. Biol. 218, 816–818 (2015).
Marchant, C. L., Malmi-Kakkada, A. N., Espina, J. A. & Barriga, E. H. Cell clusters softening triggers collective cell migration in vivo. Nat. Mater. 21, 1314–1323 (2022).
Pages, D. L. et al. Cell clusters adopt a collective amoeboid mode of migration in confined nonadhesive environments. Sci. Adv. 8, eabp8416 (2022).
Wrenn, E. D. et al. Regulation of collective metastasis by nanolumenal signaling. Cell 183, 395–410.e319 (2020).
Kim, Y. Y. et al. KDM3B inhibitors disrupt the oncogenic activity of PAX3-FOXO1 in fusion-positive rhabdomyosarcoma. Nat. Commun. 15, 1703 (2024).
Linardic, C. M. PAX3-FOXO1 fusion gene in rhabdomyosarcoma. Cancer Lett. 270, 10–18 (2008).
Zhuang, H. H., Qu, Q., Teng, X. Q., Dai, Y. H. & Qu, J. Superenhancers as master gene regulators and novel therapeutic targets in brain tumors. Exp. Mol. Med. 55, 290–303 (2023).
Li, M. et al. Dissecting super-enhancer driven transcriptional dependencies reveals novel therapeutic strategies and targets for group 3 subtype medulloblastoma. J. Exp. Clin. Cancer Res. 41, 311 (2022).
Lotem, J., Netanely, D., Domany, E. & Sachs, L. Human cancers overexpress genes that are specific to a variety of normal human tissues. Proc. Natl Acad. Sci. USA 102, 18556–18561 (2005).
Gavish, A. et al. Hallmarks of transcriptional intratumour heterogeneity across a thousand tumours. Nature 618, 598–606 (2023).
de Visser, K. E. & Joyce, J. A. The evolving tumor microenvironment: from cancer initiation to metastatic outgrowth. Cancer Cell 41, 374–403 (2023).
Gerstberger, S., Jiang, Q. & Ganesh, K. Metastasis. Cell 186, 1564–1579 (2023).
Khan, I. & Steeg, P. S. Keeping brain metastases dormant. Nat. Cancer 3, 3–5 (2022).
Khan, I. & Steeg, P. S. A perspective on the metastasis suppressor field. Cancer Metastasis Rev. 42, 1061–1063 (2023).
Khan, I. & Steeg, P. S. Metastasis suppressors: functional pathways. Lab. Invest. 98, 198–210 (2018).
Fares, J., Fares, M. Y., Khachfe, H. H., Salhab, H. A. & Fares, Y. Molecular principles of metastasis: a hallmark of cancer revisited. Signal Transduct. Target Ther. 5, 28 (2020).
Welch, D. R. & Hurst, D. R. Defining the hallmarks of metastasis. Cancer Res 79, 3011–3027 (2019).
Welch, D. R. Metastasis suppressors: a paradigm shift in cancer biology. Cancer Metastasis Rev. 42, 1057–1059 (2023).
Silies, M. et al. Glial cell migration in the eye disc. J. Neurosci. 27, 13130–13139 (2007).
Stitt, T. N., Gasser, U. E. & Hatten, M. E. Molecular mechanisms of glial-guided neuronal migration. Ann. N. Y. Acad. Sci. 633, 113–121 (1991).
Nascimento, M. A. et al. Protracted neuronal recruitment in the temporal lobes of young children. Nature 626, 1056–1065 (2024).
Francou, A., Anderson, K. V. & Hadjantonakis, A. K. A ratchet-like apical constriction drives cell ingression during the mouse gastrulation EMT. Elife 12, e84019 (2023).
Wu, J. S. et al. Plasticity of cancer cell invasion: patterns and mechanisms. Transl. Oncol. 14, 100899 (2021).
Schuster, E. et al. Better together: circulating tumor cell clustering in metastatic cancer. Trends Cancer 7, 1020–1032 (2021).
Fang, D. & Zhu, J. Dynamic balance between master transcription factors determines the fates and functions of CD4 T cell and innate lymphoid cell subsets. J. Exp. Med. 214, 1861–1876 (2017).
Li, J. et al. A single-cell chromatin accessibility dataset of human primed and naive pluripotent stem cell-derived teratoma. Sci. Data 11, 725 (2024).
Lavaud, M. et al. Super-enhancers: drivers of cells’ identities and cells’ debacles. Epigenomics 16, 681–700 (2024).
Guarnaccia, A. D. & Tansey, W. P. Moonlighting with WDR5: a cellular multitasker. J. Clin. Med. 7, 21 (2018).
Poulard, C., Noureddine, L. M., Pruvost, L. & Le Romancer, M. Structure, activity, and function of the protein lysine methyltransferase G9a. Life 11, 1082 (2021).
Heinz, S., Romanoski, C. E., Benner, C. & Glass, C. K. The selection and function of cell type-specific enhancers. Nat. Rev. Mol. Cell Biol. 16, 144–154 (2015).
Yoshino, S. & Suzuki, H. I. The molecular understanding of super-enhancer dysregulation in cancer. Nagoya J. Med. Sci. 84, 216–229 (2022).
Cappabianca, L. et al. Molecular characterization and inhibition of a novel stress-induced mitochondrial protecting role for misfolded TrkAIII in human SH-SY5Y neuroblastoma cells. Int. J. Mol. Sci. 25, 5475 (2024).
Peard, L., Ziada, A., James, A., Radulescu, V. & Saltzman, A. F. Neuroblastoma in a newborn female. Urology 185, 80–83 (2024).
Bottino, C., Vitale, C., Dondero, A. & Castriconi, R. B7-H3 in pediatric tumors: far beyond neuroblastoma. Cancers 15, 3279 (2023).
Smiles, W. J., Catalano, L., Stefan, V. E., Weber, D. D. & Kofler, B. Metabolic protein kinase signalling in neuroblastoma. Mol. Metab. 75, 101771 (2023).
Bechmann, N., Westermann, F. & Eisenhofer, G. HIF and MYC signaling in adrenal neoplasms of the neural crest: implications for pediatrics. Front. Endocrinol. 14, 1022192 (2023).
Bhardwaj, N., Das, G. & Srinivasan, R. Neuroblastoma-derived v-myc avian myelocytomatosis viral related oncogene or MYCN gene. J. Clin. Pathol. 76, 518–523 (2023).
Almeida, J. et al. Deciphering the role of p53 and TAp73 in neuroblastoma: from pathogenesis to treatment. Cancers 14, 6212 (2022).
Bobin, C. et al. Sequential analysis of cfDNA reveals clonal evolution in neuroblastoma patients receiving ALK targeted therapy. Clin. Cancer Res. https://doi.org/10.1158/1078-0432.CCR-1124-0753 (2024).
Secomandi, E. et al. Differential competitive growth of transgenic subclones of neuroblastoma cells expressing different levels of cathepsin D co-cultured in 2D and 3D in response to EGF: implications in tumor heterogeneity and metastasis. Cancers 16, 1343 (2024).
Schoof, M. et al. Mouse models of pediatric high-grade gliomas with MYCN amplification reveal intratumoral heterogeneity and lineage signatures. Nat. Commun. 14, 7717 (2023).
Hu, J. et al. Anlotinib induces neuronal-like differentiation of neuroblastoma by downregulating CRMP5. Mol. Carcinog. 63, 1559–1571 (2024).
Ekstrom, T. L. et al. USP44 overexpression drives a MYC-like gene expression program in neuroblastoma through epigenetic reprogramming. Mol. Cancer Res. https://doi.org/10.1158/1541-7786.MCR-1123-0454 (2024).
Zhang, X. et al. TAF1D promotes proliferation by transcriptionally activating G2/M phase-related genes in MYCN-amplified neuroblastoma. Cancer Sci. 114, 2860–2870 (2023).
Mei, S. et al. Single-cell analyses of metastatic bone marrow in human neuroblastoma reveals microenvironmental remodeling and metastatic signature. JCI Insight 9, e173337 (2024).
Banerjee, D. et al. Activating transcription factor 5 promotes neuroblastoma metastasis by inducing anoikis resistance. Cancer Res. Commun. 3, 2518–2530 (2023).
Liu, S. et al. Metastasis pattern and prognosis in children with neuroblastoma. World J. Surg. Oncol. 21, 130 (2023).
Heskett, C., Kabangu, J. L., Bhargav, A., Chamoun, R. & Ohiorhenuan, I. Metastatic esthesioneuroblastoma with widespread recurrence to the central nervous system. Surg. Neurol. Int 14, 77 (2023).
Xu, M. et al. HAND2 assists MYCN enhancer invasion to regulate a noradrenergic neuroblastoma phenotype. Cancer Res. 83, 686–699 (2023).
Guo, X., Chen, Q. R., Song, Y. K., Wei, J. S. & Khan, J. Exon array analysis reveals neuroblastoma tumors have distinct alternative splicing patterns according to stage and MYCN amplification status. BMC Med. Genomics 4, 35 (2011).
Maris, J. M. The biologic basis for neuroblastoma heterogeneity and risk stratification. Curr. Opin. Pediatr. 17, 7–13 (2005).
Maris, J. M., Hogarty, M. D., Bagatell, R. & Cohn, S. L. Neuroblastoma. Lancet 369, 2106–2120 (2007).
Saldana-Guerrero, I. M. et al. A human neural crest model reveals the developmental impact of neuroblastoma-associated chromosomal aberrations. Nat. Commun. 15, 3745 (2024).
Ponzoni, M. et al. Recent advances in the developmental origin of neuroblastoma: an overview. J. Exp. Clin. Cancer Res. 41, 92 (2022).
Tonini, G. P. in Neuroblastoma—Current State and Recent Updates (ed. Gowda, C.) (IntechOpen Limited, 2017).
Liu, X. et al. PCLAF promotes neuroblastoma G1/S cell cycle progression via the E2F1/PTTG1 axis. Cell Death Dis. 13, 178 (2022).
Liu, R., Shi, P., Wang, Z., Yuan, C. & Cui, H. Molecular mechanisms of MYCN dysregulation in cancers. Front. Oncol. 10, 625332 (2020).
Gryder, B., Scacheri, P. C., Ried, T. & Khan, J. Chromatin mechanisms driving cancer. Cold Spring Harb. Perspect. Biol. 14, a040956 (2022).
Liu, Z. et al. MYCN drives oncogenesis by cooperating with the histone methyltransferase G9a and the WDR5 adaptor to orchestrate global gene transcription. PLoS Biol. 22, e3002240 (2024).
Bryan, A. F. et al. WDR5 is a conserved regulator of protein synthesis gene expression. Nucleic Acids Res. 48, 2924–2941 (2020).
Koche, R. P. et al. Extrachromosomal circular DNA drives oncogenic genome remodeling in neuroblastoma. Nat. Genet. 52, 29–34 (2020).
Boeva, V. et al. Heterogeneity of neuroblastoma cell identity defined by transcriptional circuitries. Nat. Genet. 49, 1408–1413 (2017).
van Groningen, T. et al. Neuroblastoma is composed of two super-enhancer-associated differentiation states. Nat. Genet. 49, 1261–1266 (2017).
United States Environmental Protection Agency. Health—neurodevelopmental disorders. https://www.epa.gov/americaschildrenenvironment/health-neurodevelopmental-disorders (2023).
Sturm, D., Pfister, S. M. & Jones, D. T. W. Pediatric gliomas: current concepts on diagnosis, biology, and clinical management. J. Clin. Oncol. 35, 2370–2377 (2017).
Parenti, I., Rabaneda, L. G., Schoen, H. & Novarino, G. Neurodevelopmental disorders: from genetics to functional pathways. Trends Neurosci. 43, 608–621 (2020).
Kunz, L. et al. Reduced grid-cell-like representations in adults at genetic risk for Alzheimer’s disease. Science 350, 430–433 (2015).
Song, J. et al. Dysregulation of neuron differentiation in an autistic savant with exceptional memory. Mol. Brain 12, 91 (2019).
Zhang, K. et al. Imbalance of excitatory/inhibitory neuron differentiation in neurodevelopmental disorders with an NR2F1 point mutation. Cell Rep. 31, 107521 (2020).
Saunders, N. R., Habgood, M. D. & Dziegielewska, K. M. Barrier mechanisms in the brain, II. Immature brain. Clin. Exp. Pharm. Physiol. 26, 85–91 (1999).
Sturm, D. et al. Paediatric and adult glioblastoma: multiform (epi)genomic culprits emerge. Nat. Rev. Cancer 14, 92–107 (2014).
Gallo-Oller, G. et al. TRIB3 silencing promotes the downregulation of Akt pathway and PAX3-FOXO1 in high-risk rhabdomyosarcoma. Exp. Hematol. Oncol. 13, 38 (2024).
National Cancer Institute. Master regulators flip the switch on neuroblastoma’s developmental state. https://ccr.cancer.gov/news/article/master-regulators-flip-the-switch-on-neuroblastomas-developmental-state (2024).
American Cancer Society. Chemotherapy and related drugs for neuroblastoma. https://www.cancer.org/cancer/types/neuroblastoma/treating/chemotherapy.html (2023).
Zimmer, J., Rolin, C. & Ollert, M. Two hits are better than one: rational dual strategy efficiently fights neuroblastoma. Signal Transduct. Target Ther. 9, 116 (2024).
Bergaggio, E. et al. ALK inhibitors increase ALK expression and sensitize neuroblastoma cells to ALK.CAR-T cells. Cancer Cell 41, 2100–2116.e2110 (2023).
Ruella, M. ALKemy to enhance chimeric antigen receptor T cell immunotherapy for neuroblastoma. Cancer Cell 41, 2016–2018 (2023).
Qiu, B. & Matthay, K. K. Advancing therapy for neuroblastoma. Nat. Rev. Clin. Oncol. 19, 515–533 (2022).
Quamine, A. et al. TIM-3 blockade enhances ex vivo stimulated allogeneic NK cell therapy for relapsed murine neuroblastoma after hematopoietic cell transplant. bioRxiv https://doi.org/10.1101/2024.1107.1109.602731 (2024).
Murray, J. E. et al. The transcriptional co-repressor Runx1t1 is essential for MYCN-driven neuroblastoma tumorigenesis. Nat. Commun. 15, 5585 (2024).
Pathania, A. S. et al. The miR-29 family facilitates the activation of NK-cell immune responses by targeting the B7-H3 immune checkpoint in neuroblastoma. Cell Death Dis. 15, 428 (2024).
Jacobson, J. C., Qiao, J., Cochran, E. D., McCreery, S. & Chung, D. H. Migration, invasion, and metastasis are mediated by P-Rex1 in neuroblastoma. Front. Oncol. 14, 1336031 (2024).
Liu, K. X. et al. Patterns of recurrence after radiotherapy for high-risk neuroblastoma: implications for radiation dose and field. Radiother. Oncol. 198, 110384 (2024).
Vince, C. S. C. et al. Beyond clinical trials: understanding neurotrophic tropomyosin receptor kinase inhibitor challenges and efficacy in real-world pediatric oncology. JCO Precis. Oncol. 8, e2300713 (2024).
Qin, X. et al. CKLF instigates a “cold” microenvironment to promote MYCN-mediated tumor aggressiveness. Sci. Adv. 10, eadh9547 (2024).
Lukoseviciute, M. et al. Efficacy of combined targeted therapy with PI3K and CDK4/6 or PARP and WEE1 inhibitors in neuroblastoma cell lines. Oncol. Rep. 50, 166 (2023).
Broso, F. et al. Alpha-1 adrenergic antagonists Sensitize neuroblastoma to therapeutic differentiation. Cancer Res. 83, 2733–2749 (2023).
Castell, A. et al. MYCMI-7: a small MYC-binding compound that inhibits MYC: MAX interaction and tumor growth in a MYC-dependent manner. Cancer Res. Commun. 2, 182–201 (2022).
Albiach, A. M. et al. Glioblastoma is spatially organized by neurodevelopmental programs and a glial-like wound healing response. bioRxiv https://doi.org/10.1101/2023.1109.1101.555882 (2023).
Lv, X. et al. Decoding heterogeneous and coordinated tissue architecture in glioblastoma using spatial transcriptomics. iScience 27, 110064 (2024).
Hara, T. et al. Interactions between cancer cells and immune cells drive transitions to mesenchymal-like states in glioblastoma. Cancer Cell 39, 779–792.e711 (2021).
Louis, D. N., Ohgaki, H., Wiestler, O. D. & Cavenee, W. K. WHO Classification of Tumours of the Central Nervous System Revised 4th edn, Vol. 1 (IARC, 2016).
Wang, Q. et al. Tumor evolution of glioma-intrinsic gene expression subtypes associates with immunological changes in the microenvironment. Cancer Cell 32, 42–56.e46 (2017).
Azam, Z., To, S. T. & Tannous, B. A. Mesenchymal transformation: the Rosetta Stone of glioblastoma pathogenesis and therapy resistance. Adv. Sci. 7, 2002015 (2020).
Holland, E. C., Hively, W. P., DePinho, R. A. & Varmus, H. E. A constitutively active epidermal growth factor receptor cooperates with disruption of G1 cell-cycle arrest pathways to induce glioma-like lesions in mice. Genes Dev. 12, 3675–3685 (1998).
Yang, R. et al. EGFR activates GDH1 transcription to promote glutamine metabolism through MEK/ERK/ELK1 pathway in glioblastoma. Oncogene 39, 2975–2986 (2020).
Heady, D. New drug delays progression of glioma, a deadly brain cancer. https://www.uclahealth.org/news/release/new-drug-delays-progression-glioma-deadly-brain-cancer (2023).
Mellinghoff, I. K. et al. Vorasidenib in IDH1- or IDH2-mutant low-grade glioma. N. Engl. J. Med. 389, 589–601 (2023).
Bagley, S. J. et al. Intrathecal bivalent CAR T cells targeting EGFR and IL13Ralpha2 in recurrent glioblastoma: phase 1 trial interim results. Nat. Med. 30, 1320–1329 (2024).
Penn Medicine. “Dual-Target” cell therapy appears to shrink brain tumors, Penn Medicine research finds. https://www.pennmedicine.org/news/news-releases/2024/march/dual-target-cell-therapy-appears-to-shrink-brain-tumors (2024).
National Cancer Institute. Powerful drug combination kills glioblastoma tumors containing a unique genetic makeup. https://ccr.cancer.gov/neuro-oncology-branch/powerful-drug-combination-kills-glioblastoma-tumors-containing-a-unique-genetic-makeup (2023).
Drugs.com. Medications for glioblastoma multiforme. https://www.drugs.com/condition/glioblastoma-multiforme.html (2024).
Liu, J. et al. Multi-scale signaling and tumor evolution in high-grade gliomas. Cancer Cell 42, 1217–1238.e1219 (2024).
Joruiz, S. M., Von Muhlinen, N., Horikawa, I., Gilbert, M. R. & Harris, C. C. Distinct functions of wild-type and R273H mutant Delta133p53alpha differentially regulate glioblastoma aggressiveness and therapy-induced senescence. Cell Death Dis. 15, 454 (2024).
Qu, J. et al. The tumor-enriched small molecule gambogic amide suppresses glioma by targeting WDR1-dependent cytoskeleton remodeling. Signal Transduct. Target Ther. 8, 424 (2023).
Abu Rabe, D. et al. Identification of novel GANT61 analogs with activity in Hedgehog functional assays and GLI1-dependent cancer cells. Molecules 29, 3095 (2024).
Schaufelberger, S. A. et al. 2-Methoxyestradiol, an endogenous 17beta-estradiol metabolite, induces antimitogenic and apoptotic actions in oligodendroglial precursor cells and triggers endoreduplication via the p53 pathway. Cells 13, 1086 (2024).
Nagl, W. Chromatin organization and the control of gene activity. Int. Rev. Cytol. 94, 21–56 (1985).
Najem, H. et al. STING agonist 8803 reprograms the immune microenvironment and increases survival in preclinical models of glioblastoma. J. Clin. Invest. 134, e175033 (2024).
Gai, Q. J. et al. EPHA2 mediates PDGFA activity and functions together with PDGFRA as prognostic marker and therapeutic target in glioblastoma. Signal Transduct. Target Ther. 7, 33 (2022).
von Roemeling, C. A. et al. Adeno-associated virus delivered CXCL9 sensitizes glioblastoma to anti-PD-1 immune checkpoint blockade. Nat. Commun. 15, 5871 (2024).
Pearson, J. R. D. et al. TRP-2/gp100 DNA vaccine and PD-1 checkpoint blockade combination for the treatment of intracranial tumors. Cancer Immunol. Immunother. 73, 178 (2024).
Gerstner, E. R., Choi, B. D. & Curry, W. T. Infusion of CARv3-TEAM-E T cells in glioblastoma. Reply. N. Engl. J. Med. 390, 2330–2331 (2024).
Wong, E. T. Infusion of CARv3-TEAM-E T cells in glioblastoma. N. Engl. J. Med. 390, 2330 (2024).
Han, M. et al. Targeting MDM2-p53 interaction in glioblastoma: transcriptomic analysis and peptide-based inhibition strategy. Bioorg. Chem. 150, 107620 (2024).
Uemichi, Y. et al. Effects of lomerizine and its metabolite on glioblastoma cells. Anticancer Res. 44, 2943–2952 (2024).
Ter Linden, E., Abels, E. R., van Solinge, T. S., Neefjes, J. & Broekman, M. L. D. Overcoming barriers in glioblastoma-advances in drug delivery strategies. Cells 13, 998 (2024).
Tanriverdi, G. et al. The effects of the combination of temozolomide and Eribulin on T98G human glioblastoma cell line: an ultrastructural study. Ultrastruct. Pathol. 48, 323–337 (2024).
Tuncbag, N. et al. Network modeling identifies patient-specific pathways in glioblastoma. Sci. Rep. 6, 28668 (2016).
Dincer, C., Kaya, T., Keskin, O., Gursoy, A. & Tuncbag, N. 3D spatial organization and network-guided comparison of mutation profiles in Glioblastoma reveals similarities across patients. PLoS Comput. Biol. 15, e1006789 (2019).
Mathur, R. et al. Glioblastoma evolution and heterogeneity from a 3D whole-tumor perspective. Cell 187, 446–463 (2024).
Eisenbarth, D. & Wang, Y. A. Glioblastoma heterogeneity at single cell resolution. Oncogene 42, 2155–2165 (2023).
Xie, T. et al. Pervasive structural heterogeneity rewires glioblastoma chromosomes to sustain patient-specific transcriptional programs. Nat. Commun. 15, 3905 (2024).
Backx, E. et al. On the origin of pancreatic cancer: molecular tumor subtypes in perspective of exocrine cell plasticity. Cell Mol. Gastroenterol. Hepatol. 13, 1243–1253 (2022).
Luo, J. KRAS mutation in pancreatic cancer. Semin. Oncol. 48, 10–18 (2021).
Lanfredini, S., Thapa, A. & O’Neill, E. RAS in pancreatic cancer. Biochem. Soc. Trans. 47, 961–972 (2019).
Hassid, B. G. et al. Absence of pancreatic intraepithelial neoplasia predicts poor survival after resection of pancreatic cancer. Pancreas 43, 1073–1077 (2014).
Biffi, G. et al. IL1-induced JAK/STAT signaling is antagonized by TGFbeta to shape CAF heterogeneity in pancreatic ductal adenocarcinoma. Cancer Discov. 9, 282–301 (2019).
Salcedo Allende, M. T. et al. Overexpression of yes associated protein 1, an independent prognostic marker in patients with pancreatic ductal adenocarcinoma, correlated with liver metastasis and poor prognosis. Pancreas 46, 913–920 (2017).
Andricovich, J. et al. Loss of KDM6A activates super-enhancers to induce gender-specific squamous-like pancreatic cancer and confers sensitivity to BET inhibitors. Cancer Cell 33, 512–526.e518 (2018).
Revia, S. et al. Histone H3K27 demethylase KDM6A is an epigenetic gatekeeper of mTORC1 signalling in cancer. Gut 71, 1613–1628 (2022).
Waddell, N. et al. Whole genomes redefine the mutational landscape of pancreatic cancer. Nature 518, 495–501 (2015).
Sausen, M. et al. Clinical implications of genomic alterations in the tumour and circulation of pancreatic cancer patients. Nat. Commun. 6, 7686 (2015).
Lee, M. G. et al. Demethylation of H3K27 regulates polycomb recruitment and H2A ubiquitination. Science 318, 447–450 (2007).
Fritsche, P. et al. HDAC2 mediates therapeutic resistance of pancreatic cancer cells via the BH3-only protein NOXA. Gut 58, 1399–1409 (2009).
Aghdassi, A. et al. Recruitment of histone deacetylases HDAC1 and HDAC2 by the transcriptional repressor ZEB1 downregulates E-cadherin expression in pancreatic cancer. Gut 61, 439–448 (2012).
Kugel, S. et al. SIRT6 suppresses pancreatic cancer through control of Lin28b. Cell 165, 1401–1415 (2016).
Ala, M. Target c-Myc to treat pancreatic cancer. Cancer Biol. Ther. 23, 34–50 (2022).
Dhanasekaran, R. et al. The MYC oncogene—the grand orchestrator of cancer growth and immune evasion. Nat. Rev. Clin. Oncol. 19, 23–36 (2022).
Wasko, U. N. et al. Tumour-selective activity of RAS-GTP inhibition in pancreatic cancer. Nature 629, 927–936 (2024).
Ansari, D. et al. The hippo signaling pathway in pancreatic cancer. Anticancer Res. 39, 3317–3321 (2019).
Mao, W., Mai, J., Peng, H., Wan, J. & Sun, T. YAP in pancreatic cancer: oncogenic role and therapeutic strategy. Theranostics 11, 1753–1762 (2021).
National Cancer Institute. Drugs approved for pancreatic cancer. https://www.cancer.gov/about-cancer/treatment/drugs/pancreatic (2024).
Giordano, G. et al. Prognostic analysis and outcomes of metastatic pancreatic cancer patients receiving nab-paclitaxel plus gemcitabine as second or later-line treatment. Cancer Med. 13, e7345 (2024).
Lei, M. Z. et al. Acetylation promotes BCAT2 degradation to suppress BCAA catabolism and pancreatic cancer growth. Signal Transduct. Target Ther. 5, 70 (2020).
Zhang, S., Yu, Q., Li, Z., Zhao, Y. & Sun, Y. Protein neddylation and its role in health and diseases. Signal Transduct. Target Ther. 9, 85 (2024).
Rozengurt, E., Sinnett-Smith, J. & Eibl, G. Yes-associated protein (YAP) in pancreatic cancer: at the epicenter of a targetable signaling network associated with patient survival. Signal Transduct. Target Ther. 3, 11 (2018).
Wolters-Eisfeld, G., Hackert, T. & Gungor, C. Unmasking metabolic dependencies in pancreatic cancer: aberrant polyamine synthesis as a promising new therapeutic target. Signal Transduct. Target Ther. 8, 410 (2023).
De Zuani, M. et al. Single-cell and spatial transcriptomics analysis of non-small cell lung cancer. Nat. Commun. 15, 4388 (2024).
Blagosklonny, M. V. My battle with cancer. Part 1. Oncoscience 11, 1–14 (2024).
Zhu, T. et al. Mechanisms and future of non-small cell lung cancer metastasis. Front. Oncol. 10, 585284 (2020).
Riihimaki, M. et al. Metastatic sites and survival in lung cancer. Lung Cancer 86, 78–84 (2014).
Zhang, Q. et al. The spatial transcriptomic landscape of non-small cell lung cancer brain metastasis. Nat. Commun. 13, 5983 (2022).
Liu, W. et al. Protein signature for non-small cell lung cancer prognosis. Am. J. Cancer Res. 4, 256–269 (2014).
Liang, H., Xu, Y., Zhao, J., Chen, M. & Wang, M. Hippo pathway in non-small cell lung cancer: mechanisms, potential targets, and biomarkers. Cancer Gene Ther. 31, 652–666 (2024).
Choudhary, N. et al. Targeting cell signaling pathways in lung cancer by bioactive phytocompounds. Cancers 15, 3980 (2023).
Amatu, A., Sartore-Bianchi, A. & Siena, S. NTRK gene fusions as novel targets of cancer therapy across multiple tumour types. ESMO Open 1, e000023 (2016).
Vaishnavi, A., Le, A. T. & Doebele, R. C. TRKing down an old oncogene in a new era of targeted therapy. Cancer Discov. 5, 25–34 (2015).
Yan, N. et al. BRAF-mutated non-small cell lung cancer: current treatment status and future perspective. Front. Oncol. 12, 863043 (2022).
Ross, J. S. et al. The distribution of BRAF gene fusions in solid tumors and response to targeted therapy. Int. J. Cancer 138, 881–890 (2016).
Zehir, A. et al. Mutational landscape of metastatic cancer revealed from prospective clinical sequencing of 10,000 patients. Nat. Med. 23, 703–713 (2017).
Dankner, M., Rose, A. A. N., Rajkumar, S., Siegel, P. M. & Watson, I. R. Classifying BRAF alterations in cancer: new rational therapeutic strategies for actionable mutations. Oncogene 37, 3183–3199 (2018).
Konduri, K. et al. EGFR fusions as novel therapeutic targets in lung cancer. Cancer Discov. 6, 601–611 (2016).
Di Federico, A. et al. EGFR-RAD51 gene fusion NSCLC responsiveness to different generation EGFR-TKIs: two cases and review of the literature. Transl. Lung Cancer Res. 11, 497–503 (2022).
Arbour, K. C. & Riely, G. J. Diagnosis and treatment of anaplastic lymphoma kinase-positive non-small cell lung cancer. Hematol. Oncol. Clin. North Am. 31, 101–111 (2017).
Du, X., Shao, Y., Qin, H. F., Tai, Y. H. & Gao, H. J. ALK-rearrangement in non-small-cell lung cancer (NSCLC). Thorac. Cancer 9, 423–430 (2018).
Bontoux, C. et al. Deciphering the impact of HER2 alterations on non-small-cell lung cancer: from biological mechanisms to therapeutic approaches. J. Pers. Med. 12, 1651 (2022).
Pishdad, R., Illei, P. B., Gocke, C. D. & Ball, D. W. RET gene fusion and emergent Selpercatinib resistance in a calcitonin-rich neuroendocrine carcinoma: a case report. Front. Oncol. 14, 1360492 (2024).
Spitaleri, G. et al. MET in non-small-cell lung cancer (NSCLC): cross ‘a Long and Winding Road’ looking for a target. Cancers 15, 4779 (2023).
Blaquier, J. B. & Recondo, G. Non-small-cell lung cancer: how to manage MET exon 14 skipping mutant disease. Drugs Context 11, 2022-2-2 (2022).
Fernandes, M. et al. MET exon 14 skipping mutation is a hepatocyte growth factor (HGF)-dependent oncogenic driver in vitro and in humanised HGF knock-in mice. Mol. Oncol. 17, 2257–2274 (2023).
Tong, J. H. et al. MET amplification and exon 14 splice site mutation define unique molecular subgroups of non-small cell lung carcinoma with poor prognosis. Clin. Cancer Res. 22, 3048–3056 (2016).
Waarts, M. R., Stonestrom, A. J., Park, Y. C. & Levine, R. L. Targeting mutations in cancer. J. Clin. Invest. 132, e154943 (2022).
Reiter, J. G. et al. Minimal functional driver gene heterogeneity among untreated metastases. Science 361, 1033–1037 (2018).
Martinez-Marti, A. et al. Genetic evolution to tyrosine kinase inhibitory therapy in patients with EGFR-mutated non-small-cell lung cancer. Br. J. Cancer 125, 1561–1569 (2021).
Holderfield, M. et al. Concurrent inhibition of oncogenic and wild-type RAS-GTP for cancer therapy. Nature 629, 919–926 (2024).
American Cancer Society. Targeted drug therapy for non-small cell lung cancer. https://www.cancer.org/cancer/types/lung-cancer/treating-non-small-cell/targeted-therapies.html (2024).
Skoulidis, F. et al. Sotorasib for lung cancers with KRAS p.G12C mutation. N. Engl. J. Med. 384, 2371–2381 (2021).
Fakih, M. G. et al. Sotorasib plus panitumumab in refractory colorectal cancer with mutated KRAS G12C. N. Engl. J. Med. 389, 2125–2139 (2023).
Tian, Y. et al. Single-cell transcriptomic profiling reveals the tumor heterogeneity of small-cell lung cancer. Signal Transduct. Target Ther. 7, 346 (2022).
Yuan, Y. et al. miR-410 induces both epithelial-mesenchymal transition and radioresistance through activation of the PI3K/mTOR pathway in non-small cell lung cancer. Signal Transduct. Target Ther. 5, 85 (2020).
Yamaguchi, T. et al. Efficacy of chemotherapy plus immune checkpoint inhibitors in patients with non-small cell lung cancer who have rare oncogenic driver mutations: a retrospective analysis. BMC Cancer 24, 842 (2024).
Xia, H. et al. Neoadjuvant camrelizumab (an anti-PD-1 antibody) plus chemotherapy or apatinib (a VEGFR-2 inhibitor) for initially unresectable stage II-III non-small-cell lung cancer: a multicentre, two-arm, phase 2 exploratory study. Signal Transduct. Target Ther. 9, 145 (2024).
Kitai, H. et al. Combined inhibition of KRAS(G12C) and mTORC1 kinase is synergistic in non-small cell lung cancer. Nat. Commun. 15, 6076 (2024).
Rahul, E. et al. An updated account on molecular heterogeneity of acute leukemia. Am. J. Blood Res. 11, 22–43 (2021).
Saki, N., Abroun, S., Farshdousti Hagh, M. & Asgharei, F. Neoplastic bone marrow niche: hematopoietic and mesenchymal stem cells. Cell J. 13, 131–136 (2011).
Jin, W. et al. Cellular hierarchy insights reveal leukemic stem-like cells and early death risk in acute promyelocytic leukemia. Nat. Commun. 15, 1423 (2024).
Abe, S., Ishikawa, I., Harigae, H. & Sugawara, T. A new complex translocation t(5;17;15)(q11;q21;q22) in acute promyelocytic leukemia. Cancer Genet. Cytogenet. 184, 44–47 (2008).
Wang, Z. et al. MicroRNA-204 potentiates the sensitivity of acute myeloid leukemia cells to arsenic trioxide. Oncol. Res. 27, 1035–1042 (2019).
PDQ Pediatric Treatment Editorial Board. in PDQ Cancer Information Summaries (National Cancer Institute (US), 2002).
Turkalj, S., Radtke, F. A. & Vyas, P. An overview of targeted therapies in acute myeloid leukemia. Hemasphere 7, e914 (2023).
Sun, Y. et al. HOXA9 reprograms the enhancer landscape to promote leukemogenesis. Cancer Cell 34, 643–658.e645 (2018).
Yu, J. et al. Reduced H3K27me3 leads to abnormal Hox gene expression in neural tube defects. Epigenetics Chromatin 12, 76 (2019).
Collins, C. T. & Hess, J. L. Role of HOXA9 in leukemia: dysregulation, cofactors and essential targets. Oncogene 35, 1090–1098 (2016).
Das, P. & Taube, J. H. Regulating methylation at H3K27: a trick or treat for cancer cell plasticity. Cancers 12, 2792 (2020).
Takahashi, S. & Yokoyama, A. The molecular functions of common and atypical MLL fusion protein complexes. Biochim. Biophys. Acta Gene Regul. Mech. 1863, 194548 (2020).
Kuijk, E., Kranenburg, O., Cuppen, E. & Van Hoeck, A. Common anti-cancer therapies induce somatic mutations in stem cells of healthy tissue. Nat. Commun. 13, 5915 (2022).
Luo, Q. et al. Targetable leukaemia dependency on noncanonical PI3Kgamma signalling. Nature 630, 198–205 (2024).
Azizidoost, S. et al. Signaling pathways governing the behaviors of leukemia stem cells. Genes Dis. 11, 830–846 (2024).
Tan, X. et al. The dual role of FSP1 in programmed cell death: resisting ferroptosis in the cell membrane and promoting necroptosis in the nucleus of THP-1 cells. Mol. Med. 30, 102 (2024).
Ruglioni, M. et al. Understanding mechanisms of resistance to FLT3 inhibitors in adult FLT3-mutated acute myeloid leukemia to guide treatment strategy. Crit. Rev. Oncol. Hematol. 201, 104424 (2024).
Herrero Alvarez, N. et al. (89)Zr-immunoPET-guided selection of a CD33xIL15 fusion protein optimized for antitumor immune cell activation and in vivo tumour retention in acute myeloid leukaemia. Eur. J. Nucl. Med. Mol. Imaging. https://doi.org/10.1007/s00259-00024-06814-00257 (2024).
Lou, J. et al. Discovery of a covalent inhibitor that overcame resistance to venetoclax in AML cells overexpressing BFL-1. J. Med. Chem. 67, 10795–10830 (2024).
Knox, S. S. From ‘omics’ to complex disease: a systems biology approach to gene-environment interactions in cancer. Cancer Cell Int. 10, 11 (2010).
Balmain, A. Cancer as a complex genetic trait: tumor susceptibility in humans and mouse models. Cell 108, 145–152 (2002).
Grizzi, F. & Chiriva-Internati, M. Cancer: looking for simplicity and finding complexity. Cancer Cell Int. 6, 4 (2006).
Swanton, C. et al. Embracing cancer complexity: hallmarks of systemic disease. Cell 187, 1589–1616 (2024).
National Cancer Institute. BRCA gene mutations: cancer risk and genetic testing. https://www.cancer.gov/about-cancer/causes-prevention/genetics/brca-fact-sheet (2020).
Shiovitz, S. & Korde, L. A. Genetics of breast cancer: a topic in evolution. Ann. Oncol. 26, 1291–1299 (2015).
Petrucelli, N., Daly, M. B. & Pal, T. in GeneReviews (eds Adam, M. P., Feldman, J. & Mirzaa, G. M.) (University of Washington, 2023).
Breast Cancer Association, C. et al. Breast cancer risk genes—association analysis in more than 113,000 women. N. Engl. J. Med. 384, 428–439 (2021).
Beitsch, P. D. et al. Underdiagnosis of hereditary breast cancer: are genetic testing guidelines a tool or an obstacle? J. Clin. Oncol. 37, 453–460 (2019).
National Cancer Institute. Genetics of Colorectal Cancer (PDQ®)–Health Professional Version. https://www.cancer.gov/types/colorectal/hp/colorectal-genetics-pdq (2024).
Medina Pabón, M. A. & Babiker, H. M. in StatPearls (StatPearls Publishing, 2022).
Nussinov, R. & Tsai, C. J. ‘Latent drivers’ expand the cancer mutational landscape. Curr. Opin. Struct. Biol. 32, 25–32 (2015).
Nussinov, R., Jang, H., Tsai, C. J. & Cheng, F. Precision medicine review: rare driver mutations and their biophysical classification. Biophys. Rev. 11, 5–19 (2019).
Knudson, A. G. Jr. Mutation and cancer: statistical study of retinoblastoma. Proc. Natl Acad. Sci. USA 68, 820–823 (1971).
Ashley, D. J. The two “hit” and multiple “hit” theories of carcinogenesis. Br. J. Cancer 23, 313–328 (1969).
Nordling, C. O. A new theory on cancer-inducing mechanism. Br. J. Cancer 7, 68–72 (1953).
Liu, Y., Zhang, W., Jang, H. & Nussinov, R. SHP2 clinical phenotype, cancer, or RASopathies, can be predicted by mutant conformational propensities. Cell Mol. Life Sci. 81, 5 (2024).
Nussinov, R. et al. Review: Cancer and neurodevelopmental disorders: multi-scale reasoning and computational guide. Front. Cell Dev. Biol. 12, 1376639 (2024).
Liu, Y., Zhang, M., Jang, H. & Nussinov, R. The allosteric mechanism of mTOR activation can inform bitopic inhibitor optimization. Chem. Sci. 15, 1003–1017 (2024).
Jang, H., Chen, J., Iakoucheva, L. M. & Nussinov, R. Cancer and autism: how PTEN mutations degrade function at the membrane and isoform expression in the human brain. J. Mol. Biol. 435, 168354 (2023).
Yavuz, B. R. et al. Neurodevelopmental disorders and cancer networks share pathways, but differ in mechanisms, signaling strength, and outcome. NPJ Genom. Med. 8, 37 (2023).
Nussinov, R. et al. Neurodevelopmental disorders, like cancer, are connected to impaired chromatin remodelers, PI3K/mTOR, and PAK1-regulated MAPK. Biophys. Rev. 15, 163–181 (2023).
Nussinov, R., Tsai, C. J. & Jang, H. Neurodevelopmental disorders, immunity, and cancer are connected. iScience 25, 104492 (2022).
Hanahan, D. & Weinberg, R. A. The hallmarks of cancer. Cell 100, 57–70 (2000).
Hanahan, D. & Weinberg, R. A. Hallmarks of cancer: the next generation. Cell 144, 646–674 (2011).
Blagosklonny, M. V. Hallmarks of cancer and hallmarks of aging. Aging 14, 4176–4187 (2022).
Werner, H. & LeRoith, D. Hallmarks of cancer: the insulin-like growth factors perspective. Front. Oncol. 12, 1055589 (2022).
Ravi, S. et al. An update to hallmarks of cancer. Cureus 14, e24803 (2022).
Glaviano, A. et al. Mechanisms of sensitivity and resistance to CDK4/CDK6 inhibitors in hormone receptor-positive breast cancer treatment. Drug Resist. Updat. 76, 101103 (2024).
Sciarrillo, R. et al. Exploring splicing modulation as an innovative approach to combat pancreatic cancer: SF3B1 emerges as a prognostic indicator and therapeutic target. Int. J. Biol. Sci. 20, 3173–3184 (2024).
Zhang, W., Liu, Y., Jang, H. & Nussinov, R. CDK2 and CDK4: cell cycle functions evolve distinct, catalysis-competent conformations, offering drug targets. JACS Au 4, 1911–1927 (2024).
Burda, I. & Roeder, A. H. K. Stepping on the molecular brake: slowing down proliferation to allow differentiation. Dev. Cell 57, 561–563 (2022).
Nussinov, R. et al. Mechanism of activation and the rewired network: new drug design concepts. Med. Res. Rev. 42, 770–799 (2022).
Zhu, L. & Skoultchi, A. I. Coordinating cell proliferation and differentiation. Curr. Opin. Genet. Dev. 11, 91–97 (2001).
Nussinov, R., Tsai, C. J. & Jang, H. A new view of pathway-driven drug resistance in tumor proliferation. Trends Pharm. Sci. 38, 427–437 (2017).
Nussinov, R., Tsai, C. J. & Jang, H. Are parallel proliferation pathways redundant? Trends Biochem. Sci. 45, 554–563 (2020).
Ingram, K. et al. NKX2-1 controls lung cancer progression by inducing DUSP6 to dampen ERK activity. Oncogene 41, 293–300 (2022).
Tsai, C. J. & Nussinov, R. The molecular basis of targeting protein kinases in cancer therapeutics. Semin. Cancer Biol. 23, 235–242 (2013).
Rasnic, R., Linial, N. & Linial, M. Expanding cancer predisposition genes with ultra-rare cancer-exclusive human variations. Sci. Rep. 10, 13462 (2020).
Nussinov, R., Tsai, C. J. & Jang, H. Why are some driver mutations rare? Trends Pharm. Sci. 40, 919–929 (2019).
Beckman, R. A. & Loeb, L. A. Rare mutations in cancer drug resistance and implications for therapy. Clin. Pharm. Ther. 108, 437–439 (2020).
Scholl, C. & Frohling, S. Exploiting rare driver mutations for precision cancer medicine. Curr. Opin. Genet. Dev. 54, 1–6 (2019).
Baylin, S. B. & Jones, P. A. Epigenetic determinants of cancer. Cold Spring Harb. Perspect. Biol. 8, a019505 (2016).
Gray, J. S., Wani, S. A. & Campbell, M. J. Epigenomic alterations in cancer: mechanisms and therapeutic potential. Clin. Sci. 136, 473–492 (2022).
Hanahan, D. Hallmarks of cancer: new dimensions. Cancer Discov. 12, 31–46 (2022).
Nussinov, R., Tsai, C. J. & Jang, H. Protein ensembles link genotype to phenotype. PLoS Comput. Biol. 15, e1006648 (2019).
Jogi, A., Vaapil, M., Johansson, M. & Pahlman, S. Cancer cell differentiation heterogeneity and aggressive behavior in solid tumors. Ups. J. Med Sci. 117, 217–224 (2012).
Saghafinia, S. et al. Cancer cells retrace a stepwise differentiation program during malignant progression. Cancer Discov. 11, 2638–2657 (2021).
Li, Z. et al. Alveolar differentiation drives resistance to KRAS inhibition in lung adenocarcinoma. Cancer Discov. 14, 308–325 (2024).
Mayer, S. et al. The tumor microenvironment shows a hierarchy of cell-cell interactions dominated by fibroblasts. Nat. Commun. 14, 5810 (2023).
Schmitt, C. A., Wang, B. & Demaria, M. Senescence and cancer—role and therapeutic opportunities. Nat. Rev. Clin. Oncol. 19, 619–636 (2022).
Xiao, S. et al. Cellular senescence: a double-edged sword in cancer therapy. Front. Oncol. 13, 1189015 (2023).
Chen, H. A. et al. Senescence rewires microenvironment sensing to facilitate antitumor immunity. Cancer Discov. 13, 432–453 (2023).
Wang, G. et al. Possibility of inducing tumor cell senescence during therapy. Oncol. Lett. 22, 496 (2021).
Meirson, T., Gil-Henn, H. & Samson, A. O. Invasion and metastasis: the elusive hallmark of cancer. Oncogene 39, 2024–2026 (2020).
Alcolea, M. P. et al. Cancer hallmarks: piecing the puzzle together. Cancer Discov. 14, 674–682 (2024).
Paplomata, E. & O’Regan, R. The PI3K/AKT/mTOR pathway in breast cancer: targets, trials and biomarkers. Ther. Adv. Med. Oncol. 6, 154–166 (2014).
Glaviano, A. et al. PI3K/AKT/mTOR signaling transduction pathway and targeted therapies in cancer. Mol. Cancer 22, 138 (2023).
Peng, Y., Wang, Y., Zhou, C., Mei, W. & Zeng, C. PI3K/Akt/mTOR pathway and its role in cancer therapeutics: are we making headway? Front. Oncol. 12, 819128 (2022).
Wiese, W. et al. PI3K/Akt/mTOR signaling pathway in blood malignancies—new therapeutic possibilities. Cancers 15, 5297 (2023).
Dienstmann, R., Rodon, J., Serra, V. & Tabernero, J. Picking the point of inhibition: a comparative review of PI3K/AKT/mTOR pathway inhibitors. Mol. Cancer Ther. 13, 1021–1031 (2014).
Zhang, H. P. et al. PI3K/AKT/mTOR signaling pathway: an important driver and therapeutic target in triple-negative breast cancer. Breast Cancer 31, 539–551 (2024).
Dong, C., Wu, J., Chen, Y., Nie, J. & Chen, C. Activation of PI3K/AKT/mTOR pathway causes drug resistance in breast cancer. Front. Pharm. 12, 628690 (2021).
Zhang, W. & Liu, H. T. MAPK signal pathways in the regulation of cell proliferation in mammalian cells. Cell Res. 12, 9–18 (2002).
Cargnello, M. & Roux, P. P. Activation and function of the MAPKs and their substrates, the MAPK-activated protein kinases. Microbiol. Mol. Biol. Rev. 75, 50–83 (2011).
Guo, Y. J. et al. ERK/MAPK signalling pathway and tumorigenesis. Exp. Ther. Med. 19, 1997–2007 (2020).
Yuan, J., Dong, X., Yap, J. & Hu, J. The MAPK and AMPK signalings: interplay and implication in targeted cancer therapy. J. Hematol. Oncol. 13, 113 (2020).
Sun, Y. et al. Signaling pathway of MAPK/ERK in cell proliferation, differentiation, migration, senescence and apoptosis. J. Recept Signal Transduct. Res. 35, 600–604 (2015).
Gaglia, G. et al. Temporal and spatial topography of cell proliferation in cancer. Nat. Cell Biol. 24, 316–326 (2022).
Timofeev, O., Giron, P., Lawo, S., Pichler, M. & Noeparast, M. ERK pathway agonism for cancer therapy: evidence, insights, and a target discovery framework. NPJ Precis. Oncol. 8, 70 (2024).
Barbosa, R., Acevedo, L. A. & Marmorstein, R. The MEK/ERK network as a therapeutic target in human cancer. Mol. Cancer Res. 19, 361–374 (2021).
Smalley, I. & Smalley, K. S. M. ERK inhibition: a new front in the war against MAPK pathway-driven cancers? Cancer Discov. 8, 140–142 (2018).
Cheng, Y., Chen, J., Shi, Y., Fang, X. & Tang, Z. MAPK signaling pathway in oral squamous cell carcinoma: biological function and targeted therapy. Cancers 14, 4625 (2022).
Numan, Y. et al. Chemokine signaling and MAPK/ERK pathway for advanced prostate cancer treatment response: in ASCO Annual Meeting. J Clin. Oncol. 38, TPS275 (2020).
Antebi, Y. E., Nandagopal, N. & Elowitz, M. B. An operational view of intercellular signaling pathways. Curr. Opin. Syst. Biol. 1, 16–24 (2017).
Kholodenko, B. N., Rauch, N., Kolch, W. & Rukhlenko, O. S. A systematic analysis of signaling reactivation and drug resistance. Cell Rep. 35, 109157 (2021).
Nolan, A. A., Aboud, N. K., Kolch, W. & Matallanas, D. Hidden targets in RAF signalling pathways to block oncogenic RAS signalling. Genes 12, 553 (2021).
Manning, B. D. & Toker, A. AKT/PKB signaling: navigating the network. Cell 169, 381–405 (2017).
Kreileder, M., Barrett, I., Bendtsen, C., Brennan, D. & Kolch, W. Signaling dynamics regulating crosstalks between T-cell activation and immune checkpoints. Trends Cell Biol. 31, 224–235 (2021).
Tee, W. V., Tan, Z. W., Lee, K., Guarnera, E. & Berezovsky, I. N. Exploring the allosteric territory of protein function. J. Phys. Chem. B 125, 3763–3780 (2021).
Yip, H. Y. K. & Papa, A. Signaling pathways in cancer: therapeutic targets, combinatorial treatments, and new developments. Cells 10, 659 (2021).
Magits, W. et al. K128 ubiquitination constrains RAS activity by expanding its binding interface with GAP proteins. EMBO J. 43, 2862–2877 (2024).
Thines, L. et al. Disruption of Ca2+/calmodulin: KSR1 interaction lowers ERK activation. Protein Sci. 33, e4982 (2024).
Sewduth, R. N. et al. Spatial mechano-signaling regulation of GTPases through non-degradative ubiquitination. Adv. Sci. 10, e2303367 (2023).
Nussinov, R. & Jang, H. Direct K-Ras inhibitors to treat cancers: progress, new insights, and approaches to treat resistance. Annu. Rev. Pharm. Toxicol. 64, 231–253 (2024).
Yi, C. et al. Activation of orphan receptor GPR132 induces cell differentiation in acute myeloid leukemia. Cell Death Dis. 13, 1004 (2022).
Harvey, A. J. in Predictive Biomarkers in Oncology (eds Badve, S. & Kumar, G.) Ch. 12 Overview of Cell Signaling Pathways in Cancer, 167–182 (Springer, Cham, 2019).
Martin, G. S. Cell signaling and cancer. Cancer Cell 4, 167–174 (2003).
Fontebasso, Y. & Dubinett, S. M. Drug development for metastasis prevention. Crit. Rev. Oncog. 20, 449–473 (2015).
Singh, D. K. et al. 5-Azacytidine- and retinoic-acid-induced reprogramming of DCCs into dormancy suppresses metastasis via restored TGF-beta-SMAD4 signaling. Cell Rep. 42, 112560 (2023).
Ganesh, K. & Massague, J. Targeting metastatic cancer. Nat. Med. 27, 34–44 (2021).
Albert Einstein College of Medicine. How an anti-metastasis drug combination works. https://einsteinmed.edu/research-briefs/11015/how-an-anti-metastasis-drug-combination-works/ (2023).
Anderson, R. L. et al. A framework for the development of effective anti-metastatic agents. Nat. Rev. Clin. Oncol. 16, 185–204 (2019).
National Cancer Institute. Experimental cancer drug metarrestin targets metastatic tumors. https://www.cancer.gov/news-events/cancer-currents-blog/2018/metaresstin-metastatic-tumors (2018).
Schorghofer, D. et al. Late stage melanoma is hallmarked by low NLGN4X expression leading to HIF1A accumulation. Br. J. Cancer https://doi.org/10.1038/s41416-41024-02758-41419 (2024).
Liu, Y., Zhang, M., Jang, H. & Nussinov, R. Higher-order interactions of Bcr-Abl can broaden chronic myeloid leukemia (CML) drug repertoire. Protein Sci. 32, e4504 (2023).
Liu, Y., Zhang, M., Tsai, C. J., Jang, H. & Nussinov, R. Allosteric regulation of autoinhibition and activation of c-Abl. Comput. Struct. Biotechnol. J. 20, 4257–4270 (2022).
National Cancer Institute. Drugs approved for neuroblastoma. https://www.cancer.gov/about-cancer/treatment/drugs/neuroblastoma (2021).
Inman, S. FDA approves dinutuximab for high-risk neuroblastoma. https://www.onclive.com/view/fda-approves-dinutuximab-for-high-risk-neuroblastoma (2015).
National Cancer Institute. FDA approves first therapy for high-risk neuroblastoma. https://www.cancer.gov/news-events/cancer-currents-blog/2015/dinutuximab-neuroblastoma (2015).
FDA. FDA grants accelerated approval to naxitamab for high-risk neuroblastoma in bone or bone marrow. https://www.fda.gov/drugs/resources-information-approved-drugs/fda-grants-accelerated-approval-naxitamab-high-risk-neuroblastoma-bone-or-bone-marrow (2020).
Serani, S. FDA’S ODAC votes yes to eflornithine for pediatric neuroblastomas (2023).
FDA. FDA approves eflornithine for adult and pediatric patients with high-risk neuroblastoma. https://www.fda.gov/drugs/resources-information-approved-drugs/fda-approves-eflornithine-adult-and-pediatric-patients-high-risk-neuroblastoma (2023).
National Cancer Institute. Drugs approved for brain tumors. https://www.cancer.gov/about-cancer/treatment/drugs/brain (2024).
Fisher, J. P. & Adamson, D. C. Current FDA-approved therapies for high-grade malignant gliomas. Biomedicines 9, 324 (2021).
Dana-Farber Cancer Institute. FDA approves new therapy for glioma patients for first time in decades. https://www.dana-farber.org/newsroom/news-releases/2024/fda-approves-new-therapy-for-glioma-patients-for-first-time-in-decades (2024).
National Cancer Institute. Drugs approved for lung cancer. https://www.cancer.gov/about-cancer/treatment/drugs/lung#1 (2024).
National Cancer Institute. Drugs approved for leukemia. https://www.cancer.gov/about-cancer/treatment/drugs/leukemia (2024).
National Cancer Institute. Drugs approved for breast cancer. https://www.cancer.gov/about-cancer/treatment/drugs/breast (2024).
Finn, R. S. et al. Atezolizumab plus bevacizumab in unresectable hepatocellular carcinoma. N. Engl. J. Med. 382, 1894–1905 (2020).
National Cancer Institute. Drugs approved for liver cancer. https://www.cancer.gov/about-cancer/treatment/drugs/liver (2024).
American Liver Foundation. Rare forms of liver cancer. https://liverfoundation.org/resource-center/blog/rare-liver-disease-month-focus-rare-forms-of-liver-cancer/ (2021).
Canellas-Socias, A., Sancho, E. & Batlle, E. Mechanisms of metastatic colorectal cancer. Nat. Rev. Gastroenterol. Hepatol. 21, 609–625 (2024).
National Cancer Institute. Drugs approved for colon and rectal cancer. https://www.cancer.gov/about-cancer/treatment/drugs/colorectal (2024).
Antoon, R., Overdevest, N., Saleh, A. H. & Keating, A. Mesenchymal stromal cells as cancer promoters. Oncogene 43, 3545–3555 (2024).
Chen, W., Hoffmann, A. D., Liu, H. & Liu, X. Organotropism: new insights into molecular mechanisms of breast cancer metastasis. NPJ Precis. Oncol. 2, 4 (2018).
Gao, Y. et al. Metastasis organotropism: redefining the congenial soil. Dev. Cell 49, 375–391 (2019).
Carrolo, M. et al. Metastatic organotropism: a brief overview. Front. Oncol. 14, 1358786 (2024).
Axelsen, J. B., Lotem, J., Sachs, L. & Domany, E. Genes overexpressed in different human solid cancers exhibit different tissue-specific expression profiles. Proc. Natl Acad. Sci. USA 104, 13122–13127 (2007).
Windels, S. F. L., Malod-Dognin, N. & Przulj, N. Identifying cellular cancer mechanisms through pathway-driven data integration. Bioinformatics 38, 4344–4351 (2022).
Jablonski, K. P., Pirkl, M., Cevid, D., Buhlmann, P. & Beerenwinkel, N. Identifying cancer pathway dysregulations using differential causal effects. Bioinformatics 38, 1550–1559 (2022).
Liu, H. et al. CTpathway: a CrossTalk-based pathway enrichment analysis method for cancer research. Genome Med. 14, 118 (2022).
Hu, C. et al. Glioblastoma mutations alter EGFR dimer structure to prevent ligand bias. Nature 602, 518–522 (2022).
Xiong, D. et al. Structurally-informed human interactome reveals proteome-wide perturbations by disease mutations. bioRxiv https://doi.org/10.1101/2023.1104.1124.538110 (2024).
Lyu, G. Y., Yeh, Y. H., Yeh, Y. C. & Wang, Y. C. Mutation load estimation model as a predictor of the response to cancer immunotherapy. NPJ Genom. Med. 3, 12 (2018).
Mukherjee, R. et al. Regulation of PTEN translation by PI3K signaling maintains pathway homeostasis. Mol. Cell 81, 708–723.e705 (2021).
Tsai, C. J. & Nussinov, R. The free energy landscape in translational science: how can somatic mutations result in constitutive oncogenic activation? Phys. Chem. Chem. Phys. 16, 6332–6341 (2014).
Akdemir, K. C. et al. Disruption of chromatin folding domains by somatic genomic rearrangements in human cancer. Nat. Genet. 52, 294–305 (2020).
Cheng, F., Liang, H., Butte, A. J., Eng, C. & Nussinov, R. Personal mutanomes meet modern oncology drug discovery and precision health. Pharm. Rev. 71, 1–19 (2019).
Nussinov, R., Jang, H. & Tsai, C. J. The structural basis for cancer treatment decisions. Oncotarget 5, 7285–7302 (2014).
Acknowledgements
This project has been funded in whole or in part with federal funds from the National Cancer Institute, National Institutes of Health, under contract HHSN261201500003I. The content of this publication does not necessarily reflect the views or policies of the Department of Health and Human Services, nor does mention of trade names, commercial products, or organizations imply endorsement by the U.S. Government. This Research was supported [in part] by the Intramural Research Program of the NIH, National Cancer Institute, Center for Cancer Research.
Funding
Open access funding provided by the National Institutes of Health.
Author information
Authors and Affiliations
Contributions
R.N. and H.J. contributed to the study conception and design. R.N. performed the literature search and wrote original draft. H.J. performed data curation. H.J. and B.R.Y. revised the work and edited the draft. H.J. and B.R.Y. prepared the figures. All authors have read and approved the article.
Corresponding author
Ethics declarations
Competing interests
The authors declare no competing interests.
Rights and permissions
Open Access This 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/.
About this article
Cite this article
Nussinov, R., Yavuz, B.R. & Jang, H. Molecular principles underlying aggressive cancers. Sig Transduct Target Ther 10, 42 (2025). https://doi.org/10.1038/s41392-025-02129-7
Received:
Revised:
Accepted:
Published:
DOI: https://doi.org/10.1038/s41392-025-02129-7