Skip to main content
Erschienen in: Journal of Hematology & Oncology 1/2021

Open Access 01.12.2021 | Review

Metabolic landscapes in sarcomas

verfasst von: Richard Miallot, Franck Galland, Virginie Millet, Jean-Yves Blay, Philippe Naquet

Erschienen in: Journal of Hematology & Oncology | Ausgabe 1/2021

Abstract

Metabolic rewiring offers novel therapeutic opportunities in cancer. Until recently, there was scant information regarding soft tissue sarcomas, due to their heterogeneous tissue origin, histological definition and underlying genetic history. Novel large-scale genomic and metabolomics approaches are now helping stratify their physiopathology. In this review, we show how various genetic alterations skew activation pathways and orient metabolic rewiring in sarcomas. We provide an update on the contribution of newly described mechanisms of metabolic regulation. We underscore mechanisms that are relevant to sarcomagenesis or shared with other cancers. We then discuss how diverse metabolic landscapes condition the tumor microenvironment, anti-sarcoma immune responses and prognosis. Finally, we review current attempts to control sarcoma growth using metabolite-targeting drugs.
Hinweise

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Abkürzungen
3HB
3‐Hydroxybutyrate
AcCoA
Acetyl coenzyme A
AIDS
Acquired immunodeficiency syndrome
aKGDD
Alpha-ketoglutarate dioxygenase
ARMS
Alveolar rhabdomyosarcoma
AS
Angiosarcoma
ASPS
Alveolar soft part sarcoma
ATF6
Activating transcription factor 6
ATRX
α-Thalassemia/mental retardation syndrome x-linked
BCL2
B-cell cell/lymphoma 2
BCL-XL
B-cell lymphoma-extra large
BIM
BCL-2-interacting mediator of cell death
BLCA
Bladder urothelial carcinoma
BMP
Bone morphogenetic proteins
BRCA
Breast invasive carcinoma
CAF
Cancer-associated fibroblast
CDK4
Cyclin-dependent kinase 4
CDKN1A = p21
Cyclin-dependent kinase inhibitor 1a
CDKN2A = p16
Cyclin-dependent kinase inhibitor 2a
CEBPA
CCAAT/enhancer-binding protein alpha
CNV
Copy number variations
CS
Canine sarcoma
DCA
Dichloroacetate
DDIT3
DNA damage-inducible transcript 3
DDLPS
Dedifferentiated liposarcoma
DNMT1
DNA methyltransferase 1
ECM
Extracellular matrix
EF
EWS/FLI1 fusion protein
EGR1
Early growth response protein 1
EMT
Epithelial to mesenchymal transition
ERMS
Embryonal rhabdomyosarcoma
ESCA
Esophageal carcinoma
ETC
Electron transport chain
EWS
Ewing sarcoma
F2,6BP
Fructose-2,6-bisphosphate
FASL
Fas ligand
FBP
Fructose-biphosphate
FGFs
Fibroblast growth factors
FLI1
Friend leukemia integration 1 transcription factor
FRS2
Fibroblast growth factor receptor substrate 2
FS
Fibrosarcomas
FUS/DDIT3
Fused in sarcoma (FUS) to DNA damage inducible transcript 3 (DDIT3)
G6P
Glucose-6-phosphate
GFR
Growth factor receptor
GIST
Gastrointestinal stromal tumor
GLUT
Glucose transporter
GSK
Glycogen synthase kinase
GTPase
Guanosine triphosphate hydrolase
HHV8
Human herpes virus 8
HIF
Hypoxia inducible factor
HIPPO
Hippopotamus (according to the tissue overgrowth)
HK
Hexokinase
IAP
Inhibitors of apoptosis proteins
IDH1
Isocitrate dehydrogenase 1
IGF1
Insulin-like growth factor 1
IGF1R
Insulin-like growth factor 1 receptor
KRAS
Kirsten rat sarcoma viral
KS
Kaposi sarcoma
KSHV
Kaposi sarcoma-associated herpesvirus
LDH
Lactate deshydrogenase
LEF-1
Lymphoid enhancer-binding factor 1
LIHC
Liver hepatocellular carcinoma
LMS
Leiomyosarcoma
lncRNA
Long non-coding RNA
LPR5
Low-density lipoprotein receptor-related protein 5
LPS
Liposarcoma
LUAD
Lung adenocarcinoma
LUSC
Lung squamous cell carcinoma
MAPK
Mitogen-activated protein kinase
MCA
Methylcholanthrene
MCL1
Myeloid leukemia cell differentiation protein
MCT
Monocarboxylate transporter
MDH
Malate dehydrogenase
MDM2
Mouse double minute 2 homolog
MEG3
Maternally expressed 3
MEK
Mitogen-activated protein kinase
MESO
Mesothelioma
MFS
Myofibrosarcoma
MFH
Malignant fibrous histiocytoma (equivalent to UPS)
miR
Micro RNA
mitROS
Mitochondrial reactive oxygen species
MLS
Myxoid liposarcoma
MMP
Matrix metallopeptidases
MPNST
Malignant peripheral nerve sheath tumors
MSC
Mesenchymal stem cells
mtDNA
Mitochondrial DNA
mTOR
Mammalian target of rapamycin
mTORC1
Mammalian target of rapamycin complex 1
MyoD
Myoblast determination protein 1
NAD
Nicotinamide adenine dinucleotide
NADPH
Nicotinamide adenine dinucleotide phosphate
ncRNA
Non-coding RNA
NF
Neurofibromin
NFκB
Nuclear factor kappa-light-chain-enhancer of activated b
NOX
NADPH oxidase
NOXA
Phorbol-12-myristate-13-acetate-induced protein 1
OAA
Oxaloacetate
OS
Osteosarcoma
OVCA
Ovarian serous cystadenocarcinoma
OXPHOS
Oxidative phosphorylation
PARK2
Phosphatidic acid-regulated protein kinase
PAAD
Pancreatic adenocarcinoma
PDAC
Pancreatic ductal adenocarcinoma
PDC
Pyruvate decarboxylase
PDGF
Platelet-derived growth factor receptors
PDHK
Pyruvate dehydrogenase kinase
PDK
3-Phosphoinositide-dependent protein kinase
PERK
Protein kinase r-like endoplasmic reticulum kinase
PET-FDG
Fluorine-18-fluorodeoxyglucose positron emission tomography
PFK
Phosphofructokinase
PFKFB
6-Phosphofructo-2-kinase/fructose-2,6-biphosphatase
PGC1a
Peroxisome proliferator-activated receptor gamma coactivator 1-alpha
PGK1
Phosphoglycerate kinase 1
PHGDH
3-Phosphoglycerate dehydrogenase
PI3K
Phosphatidylinositol-3-kinase
PKM2
Pyruvate kinase M2
PLAUR
Plasminogen activator, urokinase receptor
PLOD2
Procollagen-lysine,2-oxoglutarate 5-dioxygenase 2
PPP
Pentose phosphate pathway
PSAT
Phosphohydroxythreonine aminotransferase
PSPH
Phosphoserine phosphatase
PTEN
Phosphatase and tensin homolog
PTPRQ
Protein tyrosine phosphatase receptor type q
RAC1
Phosphoserine phosphatase
RAS
Rat sarcoma oncogene
RASSF14
Ras-association domain family 1 isoform a
Rb1
Retinoblastoma protein
RMS
Rhabdomyosarcoma
ROS
Reactive oxygen species
RREB1
Ras-responsive element binding protein 1
S6K
S6 kinase
SCNAs
Somatic copy-number alterations
SHMT1/2
Serine hydroxymethyltransferase 1 and 2
SMARCB1
SWI/SNF-related, matrix-associated, actin-dependent regulator of chromatin, subfamily b, member 1
SS
Synovial sarcoma
STAD
Stomach adenocarcinoma
STS
Soft tissue sarcoma
SV
Structural variation
TA
Transit amplifying
TAZ
Transcriptional coactivator with PDZ-binding motif
TCA
Tricarboxylic acid cycle
TCF
T-cell factor
TEAD
Tea domain family member 1
TERT
Telomerase reverse transcriptase
TF
Transcription factor
TFAM
Transcription factor a mitochondrial
TGF-β
Transforming growth factor beta
TIGAR
p53-induced glycolysis regulatory phosphatase
P53
Tumor protein p53
TRAIL
TNF-related apoptosis-inducing ligand
TSC1/2
Tuberous sclerosis complex 1/2
TSG
Tumor suppressor gene
UPR
Unfolded protein response
UPS
Undifferentiated pleomorphic sarcoma
VEGF
Vascular endothelial growth factor
vGPCR
Viral encoded g protein-coupled receptor
VHL
Von Hippel–Lindau tumor suppressor
VNN1
Vanin-1
WDLS
Well-differentiated liposarcoma
WNT
Wingless/integrated
YAP1
Yes-associated protein
YB-1
Y-box binding protein 1

Background

Sarcomas encompass a wide variety of tumors, with more than 170 subtypes, according to the last WHO classification. They originate from the neoplastic transformation of mesenchymal cells in connective tissues [1, 2]: 87% arise from soft tissue and 13% from bone [3, 4]. Soft tissue sarcoma (STS) presents as an indolent or aggressive disease, often only diagnosed at an advanced and/or metastatic stages. Current sarcoma classification relies on histopathology that may lead to errors in up to a quarter of cases [5]. In terms of prevalence, they represent less than 1% of adult cancers, but up to one fifth of pediatric solid malignant cancers [3]. Surgery is the standard of care for patients supplemented with chemotherapy or radiotherapy [6]. Targeted therapies remain limited to tumors with well-defined oncogenic drivers [1, 2]. Clinical trials targeting immune checkpoints show low response rates, with few responsive histotypes. Finally, biomarkers or tertiary lymphoid structures may be be predictive tools for 10% of patients [7]. Consequently, improving sarcoma typing and treatment requires the use of large-scale “omics” tools to identify the oncogenic drivers, often resulting from multiple genetic alterations in adult STS. These can include translocations, mutations or amplifications/deletions that cripple major growth and differentiation pathways [812].
Given the limits of current treatments, exploiting drugs targeting metabolic pathways may pave the way to effective therapy for these largely incurable diseases.
Aggressive tumors must survive in a reorganized, stressful and metabolically competitive microenvironment. This necessary adaptation exploits tumor heterogeneity and cell networks in the tumor microenvironment. Furthermore, within a given cell, plasticity depends on interconnections between various metabolic pathways to adapt growth to the available metabolites. A major trait often amplified in these tumors is the use of aerobic glycolysis, known as the Warburg effect [13], that optimizes tumor cell growth through provision of building blocks to increase biomass [14]. Since Warburg’s discovery, a debate has existed about the persistence of mitochondrial activity in glycolytic tumors and its potential to be a drug target [15]. Despite the central role of mitochondria not only in cell energetics, homeostasis and stress sensing [16] but also reactive oxygen species (ROS) production [17] their contribution to oncogenic transformation is still debated. In some STS, germline mutations affecting mitochondrial enzymes lead to the accumulation of oncometabolites that induce a pseudo-hypoxic response and alter epigenetic marks and differentiation [18]. In the tumor microenvironment, glycolytic and oxidizing cells may compete or cooperate for an optimal use and exchange of energetic metabolites. This network involves immune cells that adapt their metabolism to exert their functions in this competitive environment [19]. The purpose of this review is to link recent findings on STS genetics to the alterations of intracellular pathways affecting their tumor metabolic landscapes. Although not necessarily specific to STS, they may represent novel therapeutic opportunities.

Unsupervised omics and single cell-based analyses highlight metabolic signatures in cancer

The development of more integrated technologies with increased sensitivity and/or resolution has helped to unravel tumor genomic and metabolic complexity in situ and to bridge the gap between mouse models and patients. Several recent studies documented the power of integrated genomic or metabolomic strategies to decipher tumors complexity. An article from The Cancer Genome Atlas (TCGA) Research Network [8] combined genetic, epigenetic and transcriptomic analyses and proposed a novel classification of STS subtypes with complex genomes. In their analysis of the number and nature of copy number variations (CNVs), they identified three dominant profiles from leiomyosarcoma (LMS), myxofibrosarcoma (MFS), undifferentiated pleomorphic sarcoma (UPS) to dedifferentiated liposarcoma (DDLPS) displaying the highest level of genomic alterations. In addition to these modifications, the nature of epigenetics marks, activating pathways or immune signatures add further prognostic value. Another article exploited TCGA data to describe the relative contribution of 114 metabolic pathways to cancer progression [20]. This analysis showed that master metabolic transcriptional regulators behave as genetic drivers explaining the metabolic profiles displayed by various tumors compared to normal tissues, and help predict responsiveness to metabolism-targeting drugs. For example, alterations of specific transcriptional regulators explain the defect in polyamine biosynthesis in prostate cancer. Similarly, distinct pathways enriched in breast cancer allow the discrimination of aggressive tumors from those associated with a good prognostic. Based on this finding, metformin, a mitochondrial complex 1 inhibitor, has been proposed as a potential adjunct therapy against basal breast cancer cells, due to its unique deregulation of the Tricarboxylic Acid (TCA) cycle. In STS, this analysis highlighted the enrichment in the pentose and glucuronate interconversion (PGI) pathway, also amplified in the Yang Huang syndrome described in the context of traditional Chinese pharmacology [21]. The PGI pathway relies on UDP-glucuronosyltransferase (UGT) enzymes that catalyze the binding of d-glucuronic acid to toxic substances or endogenous compounds such as bilirubin via glycosidic bonds, contributing to the detoxification of lipophilic compounds or glucuronides.
Exploration of the TCGA database allows one to identify more discrete signatures displayed by major STS subtypes versus other types of cancers. As shown in Fig. 1, all cancer types display abnormalities in cell cycle regulation. Most carcinomas show an enrichment in oncogenic pathways, glycolytic signatures and alterations of energetic, nucleotide, amino acid or macromolecule pathways. When considering STS as a whole, RAS, PI3K and HIPPO pathways light up, as in [8], coupled to a dominant glycolytic/OXPHOS signature. More discrete signals confirm the enhancement of the PGI pathway in STS, although this trend is not detectable when considering individually the STS subtypes. Our analysis also indicates that distinct signatures preferentially match with STS subtypes, with UPS featuring an enrichment in PPAR/fatty acids and glycine/serine/threonine pathways, whereas LMS display an enhanced OXPHOS signature. Similarly, differences in oncogenic pathway usage are apparent but it is difficult to relate these pathways to the metabolic bias in tumors.
The improvement of chromatographic and mass spectrometry (MS) analyses such as Ultra High Performance Liquid Chromatography Q-Exactive MS (UHPLC-QE MS) has allowed time to be saved in sample separation while preserving the detection capacity of a large spectrum of metabolites. In osteosarcoma (OS), studies combining state-of-the-art transcriptomic and metabolomics approaches highlighted the amplification of nucleotide and amino acid (namely alanine, aspartate, glutamate, arginine, proline, methionine) pathways, glycolysis and the pentose phosphate shunt [26, 27]. Spatially-correlated analysis, mass spectrometry imaging (MALDI-MSI) can further reveal how biomolecular ions are distributed on tissue sections, linking their molecular identification to their spatial distribution. Results from two studies [28, 29] comparing STS subtypes showed that the overexpression of acyl-CoA-binding protein and stearyl-CoA desaturase (directly involved in the processing of fatty acids) as well as MIF1, galectin 1, thioredoxin could help distinguish LMS from MFS, and predict their prognostic. To identify tumor-associated metabolites in situ, airflow-assisted desorption electrospray ionization mass spectrometry imaging (AFADESI-MSI) was used on tissues from 256 esophageal cancer (ESCA) patients [30]. This analysis unraveled the dysregulation of several metabolic pathways affecting proline, glutamine, histidine, uridine, fatty acid and polyamine homeostasis. Among others, it identified pyrroline-5-carboxylate reductase 2 (PYCR2) and ornithine decarboxylase (ODC), rate-limiting enzymes in proline and polyamine biosynthesis, respectively, as markers of tumor proliferation. Table 1 summarizes biomarkers and metabolites linked to alterations in the activation of metabolic pathways in STS. In the future, these explorations will benefit from single-cell strategies evaluating the metabolic status of tumor versus surrounding cells. In this context, the development of novel flow cytometry-based methods to assess metabolic activity such as Met-Flow [31] or SCENITH [32] have already proven their potential to assay the metabolic status of circulating or tumor-infiltrating immunocytes.
Table 1
Biomarkers and metabolites associated with STS
 
Cell processes
Biomarkers (genes or metabolites)
STS subtype
Prognosis significance
References
Signaling
RAS signaling
GLUT, HK, PFK
UPS, MFS
Poor prognosis
[33, 34]
PI3K-AKT signaling
 
LMS, EWS
Poor prognosis
[35, 36]
miR-181b
STLMS, ULMS
RFS
[8]
MDM2 amplification
DDLPS
Poor prognosis
 
GFR signaling
IGFR1 overexpression
STLMS, EWS, MLS, ARMS, SS
Poor RFS/DSS
[8, 37, 38]
Her4/Erbb4
OS, EWS
Poor prognosis
[38]
Serum bFGF, VEGF
STS
Poor prognosis
[39]
JUN signaling
 
DDLPS
Poor prognosis
[8]
HIPPO pathway
Nuclear YAP/TAZ, VGLL3
UPS, MFS, MLS, RMS
Poor prognosis
[4045]
WNT pathway
Nuclear β-catenin/LEF1; MEG3 (lncRNA) downregulation
EWS, OS
Poor prognosis
[46, 47]
Cell cycle/death
Cell cycle
CINSARC—67 genes
STS
Poor prognosis
[48]
TP53, RB1, CDKN2A deficiency
LMS, UPS, MFS
Poor prognosis
[8, 4951]
TP53, IGAR, GLUT
LMS, UPS, MFS, EWS
Poor prognosis
CDCA2, KIF14, IGBP7
SS
Metastasis
[52]
CDK4 amplification
DDLPS
Poor prognosis
[8]
DNA replication
TOP2A
MPNST
Poor prognosis
[8]
RRM1
OS, EWS
Good prognosis
[8]
ATRX deletion
DDLPS
Poor prognosis
[8]
Transcriptional regulation
DNA hypermethylation,
DDLPS, STLMS
poor RFS/DSS
[8]
HMGA2 amplification
DDLPS
Poor prognosis
[8]
Energetic pathways
Glycolysis
GLUT, ENO1, TPI1, PKG1, LDHC, lactate, pyruvate
STS
Poor prognosis
[20, 53, 54]
FBP2 loss
LPS, LMS, FS, UPS
Poor DSS
[55]
Serum LDH
ULMS, EWS
Poor prognosis
[39, 51]
PKM1/2 isoenzymes
OS
Poor prognosis
[56]
Pentose and glucoronate interconversions
UGT
STS
Prognosis
[20]
Citrate cycle/OXPHOS
Downregulated metabolites
OS
Poor prognosis
[27]
Decreased ATP Synthase subunits
OS
Poor prognosis
[56]
SDH, FH mutations (succinate accumulation)
GIST
Poor prognosis
[57]
Others
AMPKa, CHK1, S6, ARID1A, RBM15, MSH6, AcetylTubulin
STS
Combined survival related signature
[58]
 
Nucleotide metabolism
 
STS
Poor prognosis
[57]
Amino acids
Alanine, aspartate, glutamate
GLS
OS, KS, EWS
High risk
[27, 59, 60, 56]
Arginine, ornithine
ASS1 deficiency, ODC
OS, MFS, KS
DSS, MFS
[27, 61, 62]
Proline
PYCR2
OS, KS
DSS, MFS
[27, 63]
Serine, glycine
PHGDH, PSAT1, PSPH, SHMT2, SLC1A5, MTHFD2, MTHFD1L
EWS
DSS, MFS
[60, 64]
Tryptophane
TDO2 (low)
EWS
DSS, MFS
[65]
5 methylthioadenosine
 
OS
DSS, MFS
[27]
Redox, vitamins
Pantothenate metabolism
VNN1 (low)
FS
Poor prognosis
[66]
Redox metabolism
TXR, MIF1, GAL1, AcCoaBP
LMS (high), MFS (low)
Poor prognosis
[67]
Hypoxia
HIF1α, hypoxia gene signatures
EWS, OS, GIST, KS
Poor prognosis
[6872]
Here, we will review and update the mechanisms that link complex oncogenic stimuli associated with various STS to metabolic alterations.

Oncogenic drivers upstream of growth pathways rewire metabolism in sarcomas

Physiologically, growth factor receptors (GFR) trigger the RAS/MAPK and PI3K/AKT/mTOR pathways and activate transcriptional regulators such as JUN/FOS/EGR1 that drive cell division (Fig. 2A). This process is temporally regulated by the co-engagement of restriction points controlled by tumor suppressor genes (TSGs) (Fig. 2B) [73]. In cancer cells, prolonged exposure to oncogenic signals strongly stimulates ERK-dependent EGR1 activation, bypassing cell cycle regulation and provoking PI3K activation that antagonizes p53-dependent tumor suppression [74] (Fig. 2B). In p53-mutated cancers, the temporal regulation of MEK/MYC/PI3K is dysfunctional and this allows cancer cells exposed to transient growth signals to proliferate, in a context of increased genomic instability. These pathways have been shown to be involved in sarcomagenesis in both human and rodent models (Fig. 2A), downstream of oncogenic GFRs or receptors involved in tissue organization and trophicity. We will highlight how various activation pathways engage these metabolic programs to sustain STS cell growth and rely alternatively on various carbon sources such as glucose, amino acids or lipids (Fig. 3). We also provide an update on current clinical trials exploiting metabolic interference in Table 2.

Overactivated MAP and PI3 kinase pathways drive a Warburg effect

Both mutations and oncogene-driven overexpression of GFRs contribute to STS development. Gain-of-function mutations of the GFR KIT or PDGF-Rα drive gastrointestinal stromal tumor (GIST) progression [93]. In the translocation-associated Ewing’s sarcoma (EWS), myxoid liposarcomas (MLS) or alveolar rhabdomyosarcoma (ARMS), the fusion proteins EF, FUS-DDIT3 or PAX3-FOXO1, respectively, enhance IGF1R expression, a major driver of RAS/AKT/mTOR activation [37]. Hyperactivation of the RAS pathway is predictive of a high risk of disease recurrence and impaired overall survival in 30% undifferentiated pleomorphic sarcoma (UPS), a common adult STS [33, 34]. Similarly, the loss of the phosphatase PTEN induces growth-factor independent PI3K/AKT activation that sustains autonomous nutrient uptake in some LMS or MPNST [94]. Accordingly, sarcoma incidence increases in hereditary neurofibromatosis patients with carrying deletions of the RAS negative regulator genes NF1 or NF2 [49, 95]. To investigate the mechanisms of tumor progression in a mouse model, whole-exome sequencing was performed on STS induced by either KrasG12D activation/p53 deletion, 3-methylcholanthrene (MCA) or ionizing radiation [96]. Whereas CNVs were very frequent in radiation-induced STS, MCA-induced tumors showed a high mutational burden, combined with high genomic instability in the absence of p53. Candidate oncogenic drivers affecting MAPK signaling were identified either as mutations of Kras, Nf1 and Hippo effectors (Fat1/4), or as amplification of Kras and Myc in p53 deficient mice, or Met and Yap1 in radiation-induced STS. Mutations in the RAS pathway influence the prognosis of human STS such as DDLPS or pediatric embryonic rhabdomyosarcoma (ERMS) [97]. In the latter, the overactivation of p38 MAPK induces high levels of reactive oxygen species (ROS) that increase the mutation rate [97] and may sensitize tumors to therapies enhancing oxidative stress [98, 99].
RAS- or PI3K/AKT-driven activation increases the expression of glucose importers (GLUT) and of the upstream ATP-consuming glycolytic enzymes hexokinase (HK) and phosphofructokinase (PFK), recently shown to control the glycolytic flow quantitatively [100]. The oncogenic KRAS4A isoform and to lesser extent other RAS isoforms were shown to interact with HK1 on mitochondria (Fig. 3), preventing its allosteric inhibition by glucose-6-phosphate (G6P), thereby enhancing glycolysis [101]. Hypoxia or mutations affecting the RAS pathway modulated the activity of PKM2 or PGK1, two ATP-generating enzymes in the last steps of glycolysis, thus providing them with non-metabolic pro-oncogenic functions [102]. Thus, an increase in the proportion of PKM2 dimers, lacking pyruvate kinase activity, drives PKM2 nuclear translocation where it participates in STAT3 phosphorylation and mek5 gene transcription, driving cell growth [103]. Another study showed that activated ERK phosphorylates PGK1, promoting its association with PIN1 and its translocation into the mitochondria. There, it phosphorylates and activates pyruvate dehydrogenase kinase 1 (PDK1), an inhibitor of pyruvate dehydrogenase (PDH), the checkpoint of pyruvate entry in the TCA cycle [104] (Fig. 3). Globally, these RAS-driven effects reinforce glycolysis over mitochondrial respiration and favor glucose- and glutamine-dependent anabolism as shown in a pancreatic ductal adenocarcinoma (PDAC) model [105, 106]. Several clinical trials are currently based on drugs inhibiting PI3K, AKT, mTOR and ERK signaling in STS (Fig. 4 and Table 2).
Table 2
Clinical trials affecting metabolic pathways in STS
 
Biomarker target
Therapeutic agent
Tumor type
Biomarker relevance/clinical trial phase
N° Clinical trial
References
MAPK pathways
RAF
Dabrafenib
Advanced solid tumors with BRAF mutations
Phase II
NCT02465060
[75]
[A]
[B]
Vemurafenib
Relapsed or refractory advanced solid tumors with BRAF V600 mutations
Phase II
NCT03220035
Dabrafenib + trametinib
MULTISARC
Phase III
NCT03784014
Dabrafenib + trametinib
BRAF V600E- mutated rare cancers
Phase II
NCT02034110
MEK1/2
Binimetinib + pexidartinib
Advanced GIST
Phase I completed
NCT03158103
[A]
[B]
[C]
Trametinib
Advanced solid tumors with BRAF mutations
Phase II
NCT02465060
Cobimetinib + MPDL3280A
Locally advanced or metastatic solid tumors
Phase I
NCT01988896
GDC-0941 + GDC-0973
Locally advanced or metastatic solid tumors
Phase II
NCT00996892
ERK1/2
Ulixertinib
STS, OS, EWS
Phase I/II
NCT03520075
[C]
PI3K/AKT/mTOR signaling
PIK3CA/mTOR
Samotolisib
STS
GIST
Phase I/II
NCT02008019
[C]
[76]
Pediatric sarcoma
Phase II
MATCH trial
NCT03458728
NCI MATCH EAY131-Z1F
[77]
[78]
GDC-0941
Locally advanced or metastatic solid tumors
Phase I
NCT00876109
[A]
GDC-0980
Locally advanced or metastatic solid tumors
Refractory solid tumors
Phase I
NCT00876122NCT00854152
NCT00854126
[A]
AKT/ERK
ONC201
Desmoplastic small round cell tumor
In vitro
 
[C]
GDC-0973 + GDC-0068
Locally advanced or metastatic solid tumors
Phase I
NCT01562275
[A]
mTOR
Sirolimus + pexidartinib
STS MPNST
Phase I/II
NCT02584647
[79]
[80]
[A]
Rapamycin + gemcitabine
OS
Phase II completed
NCT02429973
nanoparticle albumin-bound rapamycin + pazopanib
Advanced nonadipocytic soft tissue sarcomas
Phase I/II trial
NCT03660930
Lenvatinib + everolimus
Refractory pediatric solid tumors
Phase I/II
NCT03245151
CCI-779
STS/GIST
Phase II
NCT00087074
Cixutumumab + temsirolimus
Locally advanced, metastatic, or recurrent STS or bone sarcoma
Phase II
NCT01016015
CP-751,871 + RAD001
Advanced sarcomas and other malignant neoplasms
Phase I
NCT00927966
Everolimus
RAD001/progressive sarcoma
Phase II
NCT00767819
HIPPO
YAP/TAZ
Verteporfin
High histological grade
Reduced EWS metastatic potential
 
[81]
[82]
[83]
[84]
TCA CYCLE
IDH 1
IDH 1—AG-120
Chondrosarcoma
Phase I
NCT02073994
[85]
[86]
[87]
IDH 1—FT-2102
Advanced solid tumors
Active
NCT03684811
IDH 1—IDH305
Advanced malignancies with IDH1R132 mutations
Phase I
NCT02381886
IDH 1—BAY1436032
IDH1-mutant advanced solid tumors
Active
NCT02746081
AG-881
Advanced solid tumors with an IDH1 and/or IDH2 mutation
Phase I
NCT02481154
AG-120 + nivolumab
IDH1 mutant tumors
Phase II
NCT04056910
IDH2
IDH 2—AG-221
Advanced solid tumors
Phase I/II
NCT02273739
TCA cycle enzymes
Devimistat
STS
FDA orphan drug designation
 
[D]
Amino acids
ASS1 deficiency
ADI-PEG20 + gemcitabine + docetaxel
STS, OS, EWS
Phase II
NCT03449901
[88]
[89]
PDK
DCA
FS
Mice
 
[90]
GLS
CB-839—glutaminase inhibitor
GIST
Phase I completed
NCT02071862
[59]
Telaglenastat
NF1 mutation positive MPNST
Phase II
NCT03872427
Telaglenastat
 + talazoparib
Solid tumors
Phase I + phase II
NCT03875313
Heparan sulfate proteoglycans
Sulfen
EWS
Zebrafish model
 
[91]
NAMPT
FK866—MV87 inhibitors
FS
Mice
 
[92]
Folate receptor α
Pemetrexed
STS
Phase II
NCT04605770
[C]
Lipid metabolism
CPI-613
CCS
Phase II
NCT01832857
[A]
[E]
[A] The Life Raft Group. Gisttrials. https://​gisttrials.​org/​iLRG/​showfirstline.​php. Accessed 16 June 2021
[B] NIH U.S. National Library of Medicine. Clinicalstrial.gov. https://​clinicaltrials.​gov/​ct2/​home. Accessed 16 June 2021
[D] Rafael Pharmaceuticals, Inc. https://​rafaelpharma.​com/​research-and-development/​cpi-613-drug. Accessed 16 June 2021
[E] ICH GCP. Good Clinical Practice Network. https://​ichgcp.​net/​clinical-trials-registry/​NCT04593758. Accessed 16 June 2021

An altered HIPPO pathway induces aerobic glycolysis in STS

Several sarcoma histiotypes re-express genes involved in developmental pathways [107110], such as HIPPO that controls organ size. Its engagement in intercellular adhesion complexes activates the MST and LATS kinases that phosphorylate the transcriptional factors YAP1 and TAZ, promoting their degradation. In contrast, upon nuclear translocation, YAP/TAZ cooperate with mitogenic effectors and boost proliferation (Fig. 2A). HIPPO interferes with the RAS and PI3K pathways that control cell death induction [111], acting as a tumor suppressor pathway [112]. Through cross-inhibition, MST and AKT differentially regulate the expression level of pro-apoptotic effectors (NOXA, FASL, BIM, TRAIL). In addition, ERK induces the expression of anti-apoptotic effectors (BCL2, BCL-XL, IAP, MCL1) partly via YAP/TAZ activation. Furthermore, in a PDAC model, YAP1 amplification can bypass the need for oncogenic KRAS activation [113]. The amount of translocated YAP/TAZ determines their co-activating potential for TEAD transcriptions factors and thereby the balance of proliferation versus cell death [40].
Loss of MST/LATS or YAP overexpression is pro-tumoral in mice [114, 115], and this pathway is often affected in MCA- or radiation-induced STS [96] and UPS [8]. In transgenic models, altering HIPPO effectors alone or in combination with other deficits augmented STS frequency [41]. An increased YAP/TAZ nuclear staining is predictive of poor survival in UPS, DDLPS and ERMS [4245, 116]. Two studies investigated how fusion gene products mediate sarcomagenesis through alteration of the HIPPO pathway in mice. One study explored MLS that account for 5–10% STS, among which 90% tumors depend on the product of the FUS:DDIT3 translocation [44]. The authors performed a large-scale RNA interference screen and identified YAP1 as a non-redundant oncogenic driver. In MLS cell lines, FUS:DDIT3 led to a two to threefold increase in expression and co-transcriptional activity of YAP1. Co-immunoprecipitation and immunofluorescence studies revealed a physical interaction of YAP1 with FUS:DDIT3 in the nucleus. Pharmacological inhibition of YAP1 activity inhibited the growth of MLS xenografts. The other study, based on a new transgenic model, showed that doxycycline (DOX)-induced expression of YAP1 in the myogenic MYOD1 cell lineage provoked the development of ERMS through the transformation of activated satellite cells [45]. Retrieval of DOX released a YAP1-dependent differentiation block and reduced tumor formation. Transcriptional profiling of the tumors revealed that YAP1 induces pro-oncogenic effector genes and represses terminal differentiation of myoblasts. In line with these observations, an independent study found no evidence that mutant RAS isoforms were responsible for YAP overexpression in myoblasts [116]. In ARMS, the translocation product PAX3-FOXO1 suppresses HIPPO signaling through overexpression of RASSF4, which inhibits the MST1 kinase. Similarly, this effect is linked to PAX3-FOXO1 co-localization with YAP1 in the nuclei of cancer cells. This chimeric transcription factor cooperates with YAP1/TEAD to induce downstream effectors that trigger IGFs and NF-κB activation, and repress senescence and apoptosis in mesenchymal cells [40, 44]. Therefore, mutations of HIPPO effectors can be oncogenic in STS.
The HIPPO pathway restricts tissue growth and is connected to nutrient cues [117, 118]. Upon glucose starvation, AMPK- and LATS-kinases phosphorylate YAP resulting in its degradation [119]. Furthermore, Wang et al. showed that the phosphorylation level of YAP1 at position S61 is regulated by AMPK, itself recruited to YAP protein complexes in the cytosol of glucose-deprived cells. Addition of glucose was associated with a decrease in YAP phosphorylation and its nuclear translocation, where it interacted with TEAD transcriptional regulators to induce glycolytic genes. This glucose-sensing pathway via YAP and TAZ was required for the full deployment of glucose growth-promoting activity in breast cancer. In addition, glycolysis was required to sustain YAP/TAZ tumorigenic properties [120]. Mechanistically, phosphofructokinase (PFK1) bound to and co-activated the YAP/TAZ transcriptional cofactors TEADs (Fig. 3). In some cancers, the loss of NF2, an upstream negative regulator of HIPPO signaling, simultaneously unleashed YAP/TAZ and SMAD2/3 activation leading indirectly to the induction of aerobic glycolysis via derepression of GLUT, HK2, LDH and MCT genes [121]. Interestingly, NF2 mutations might contribute to the maintenance of rare aggressive sarcomas [122]. This hypothesis was tested in a kidney cancer cell model bearing NF2 mutations [123]. There, DOX-induced expression of shRNA downregulating YAP/TAZ expression provoked the regression of tumors in vivo. YAP/TAZ-depletion induced a substantial decrease in EGFR and AKT phosphorylation, associated with a reduction in glucose uptake, and a switch to glutamine anaplerosis that boosted mitochondrial respiration and ROS production. Under conditions of glucose or glutamine withdrawal, this metabolic shift favored cell death. Whereas restoration of AKT signaling by expression of a constitutively active form of AKT rescued cell proliferation, it did not prevent starvation-induced death. In vivo, YAP/TAZlow tumors survived due to the engagement of a compensatory lysosome-mediated activation of MAPK signaling. By combining YAP/TAZ and MEK inhibition, tumor growth durably regressed. YAP can also induce aerobic glycolysis through a direct interaction with HIF1α in a hypoxic environment [124, 125]. Therefore, the proglycolytic effect of YAP/TAZ engagement depends on their participation in various nuclear transcriptional complexes. In a muscle-derived UPS model, a combination of epigenetic modulators suppressed YAP1 activity and reduced sarcomagenesis through regulation of metabolism. In this case, YAP1 nuclear translocation was associated with a poor prognosis [126] and its inactivation by epigenetic modulators allowed the restoration of a clock gene-mediated unfolded protein response and muscle differentiation. It also promoted a switch toward lipid catabolism and autophagy, limiting YAP-driven UPS cell growth. The YAP/TAZ inhibitor Verteporfin is currently being tested in Ewing’s sarcoma (EWS) (Fig. 4 and Table 2).

Glutamine and arginine metabolic pathways contribute to STS growth

In a UPS mouse model harboring Kras mutations and p53 deletion in the muscle [59], tumors developed in the hindlimb and metastasized in the lung, as in the human disease. Additional deletion of HIF-2α or its binding partner aryl hydrocarbon receptor nuclear translocator (ARNT) enhanced tumor development. Use of an unbiased pan metabolomics strategy combining LC–MS and stable isotope metabolite tracing revealed a reliance on glutaminolysis for tumors, unlike muscle cells. Accordingly, glutaminase (GLS) inhibitors blocked UPS tumor growth in vivo and are the object of clinical trials in humans (Fig. 4 and Table 2). Mechanistically, GLS hydrolyses glutamine to glutamate, which is then dehydrogenated to alpha-ketoglutarate (αKG) by glutamate dehydrogenase GLUD or an aminotransferase (such as PSAT), boosting mitochondrial anaplerosis [127]. In this UPS model, tracing using C13- or N15-labeled glutamine demonstrated that glutamine is a carbon donor for the TCA cycle and a nitrogen donor for aspartate production from oxaloacetate. Aspartate is a crucial carbon source for purine and pyrimidine synthesis and sustains cell growth [128, 129]. Aspartate is also required for the conversion of citrullin into arginine through the activity of the rate-limiting argininosuccinate synthase 1 (ASS1) that initiates the urea cycle. This reaction generates arginine and contributes to the clearance of nitrogenous wastes. ASS1 deficiency has been observed in various cancers, including MFS, due to epigenetic silencing of its promoter [61]. Reexpression of ASS1 inhibited tumor growth and metastases. To investigate how arginine auxotrophy induced by ASS1 deficiency contributed to the progression of tumors, another study used a pegylated arginine deiminase (ADI-PEG20) to deplete arginine pharmacologically [89]. A short-term treatment by ADI-PEG20 applied to LMS cell lines, immediately induced cell proliferation arrest and autophagy. Upon prolonged therapy, cell lines became resistant to ADI-PEG20 due to the reexpression of ASS1 that regenerated arginine. A metabolomic profiling of treated cell lines revealed a reduction in PKM2 levels. In addition, U13C glucose tracing studies showed that carbons were shifted away from lactate and citrate production, and reoriented toward serine/glycine synthesis. Analysis of metabolic requirements for growth showed a reduced reliance on glucose and a reinforcement in OXPHOS and glutaminolysis, as an alternative source of TCA cycle intermediates via anaplerosis. In STS, a clinical trial using ADI-PEG20 in combination with Gemcitabine and Docetaxel has been launched (Fig. 4 and Table 2). Similarly, targeting glutamine metabolism through GLS inhibition could provoke the lethality of ASS1-deficient cancers and is currently being evaluated in GIST and NF1-mutated cancers (Fig. 4 and Table 2).

Complex metabolic rewiring in Kaposi’s sarcoma

Kaposi’s sarcoma (KS) is caused by a lytic oncogenic herpes virus (KSHV/HHV8), infecting endothelial cell precursors in immunosuppressed individuals. Infection, which is necessary but not sufficient for the growth of KS lesions, leads to the development of a vascular neoplasm associated with cytokine dysregulation driven by the virally encoded G protein coupled receptor (vGPCR). In lytically infected cells, vGPCR induced Rac1/NOX-dependent production of ROS that activated the redox sensitive STAT3 and HIF pathways [130]. Infected cells had increased lactate production and decreased mitochondrial respiration, a phenotype in part attributable to HIF1α activation [131]. Indeed, aerobic glycolysis favored by PKM2 induction sustains the maintenance of KS cells [132] (Fig. 3). Infected cells also exert a paracrine effect on neighboring endothelial cells. Indeed, PKM2 acts as a coactivator of HIF1α reinforcing the production of angiogenic cytokines [131]. Among them, PDGFRA, found phosphorylated in KS-biopsies [133], plays a major oncogenic role. HIF1α also participates in the reactivation of latently infected cells [134]. Upon transformation by KHSV, however, endothelial cells depended on glutamine for proliferation. KHSV also provoked an increase in ASS1 expression, in part through the action of KHSV-encoded miRNAs [62], leading to increased arginine production. Knockdown of ASS1 inhibited cell proliferation and iNOS-dependent, arginine-derived NO production. Treatment of KS cells with a NO donor-activated STAT3 without affecting ROS cell levels. A recent article questioned the relevance of these metabolic changes by comparing 2D versus 3D cultures of KHSV-infected cells [135]. An unbiased metabolomics analysis revealed significant changes in the levels of various non-essential amino acids in 3D cultures. GST-pull down studies showed that the viral K1 protein physically interacted with and activated the pyrroline-5-carboxylate reductase PYCR leading to increased proline production. This phenotype, abrogated by PYCR depletion, promoted 3D spheroid culture and tumorigenesis in nude mice. These results highlight the complex metabolic rewiring that occurs during infection and transformation by KHSV but also the need for appropriate in vitro culture systems to evaluate metabolic adaptation.

One carbon metabolism is overactive in aggressive STS

In Ewing’s sarcoma (EWS), the fusion protein resulting from a single translocation event between the regulatory domain of EWS and the DNA-binding domain of FLI1 behaves as a chimeric transcription factor called EF that enhances IGFR1 activation (Fig. 2A). Transcriptomic studies [60, 64] identified EF as an upstream regulator of PHGDH, PSAT, PSPH and SHMT1/2 genes involved in serine-glycine biosynthesis as well as SLC1A4/5 glutamine transporter genes (Fig. 3). Accordingly, knockdown of EF reduced the proportion of glucose-derived 3-phosphoglycerate reoriented toward serine and glycine synthesis; EWS cell lines were highly dependent on glutamine for growth and survival. Earlier work using a metabolomics approach with isotope labeling had already shown that a large proportion of glycolytic carbon was diverted into serine and glycine metabolism in melanoma; this was due to the amplification of the PHGDH gene [136], also elevated in high-risk EWS patients [100]. Serine or glycine can provide one carbon to tetrahydrofolate initiating the folate cycle. The enhancement of one-carbon metabolism, considered as an integrator of nutrient status [97], boosts the interconnected folate and methionine cycles leading to enhanced NADPH and nucleotide synthesis. NADPH regulates ROS-dependent death and methyl transfer contributes to epigenetic modifications. Knockdown of PHGDH recapitulated the effect of anti-metabolite chemotherapies and had a major effect on cell growth and epigenetic control. Two studies investigated the sensitizing effect of methionine restriction on chemo- or radio resistant models of RAS-driven colorectal cancer and STS, respectively [137, 138]. In the FSF KrasG12D/+; Tp53−/− STS mouse model, tumor development was triggered by the intramuscular injection of an adenovirus carrying the FlpO recombinase. In these aggressive tumors, only the combination of diet and radiation delayed tumor growth. By combining tumor metabolomics and metabolite tracing with a time-course analysis of data, alterations were observed for nucleotide and redox metabolisms. Interestingly, the consequences of methionine restriction could be detected at the metabolic level when applied to healthy individuals. These results indicated that a targeted dietary manipulation could improve tumor response to therapies.

Linking Wnt signaling alterations to metabolic rewiring

Physiologically, the canonical Wnt pathway participates to the maintenance of stem cell pools and cell fate, in part via the nuclear translocation of β-catenin, leading to its interaction with TCF/LEF transcription factors [139]. In a model of osteoblast differentiation, Wnt3a signaling induces aerobic glycolysis by increasing the level of glycolytic effectors (HK2, LDHA, PDK1, GLUT1). This process requires LRP5-mediated mTORC2/AKT activation but not β-catenin [140]. Other studies showed that Wnt engagement also reduced nuclear acetyl coenzyme A (AcCoA) levels and consequently impaired osteoblastic gene expression [141]. In contrast, in mature osteoblasts, Wnt-LRP5 boosted fatty acid oxidation and was required for bone mass increase [142]. In osteosarcoma (OS), Wnt participates in bone remodeling, maintenance of stem cell niches and EMT in collaboration with TGF-β and BMP signaling (reviewed in [143, 144]). In EWS, synovial sarcoma (SS), OS, and to a lesser extent in LMS [143, 145], a high level of Wnt activation, scored by the nuclear localization of β-catenin or LEF-1, is associated with a poor clinical outcome [46]. In OS, deletion of Wnt-related genes has been reported [146]. In some primary OS, the loss of the tumor suppressor RASSF1A enhanced Wnt activation through the AKT/GSK-3-Wnt/β-catenin pathway [147]. Other studies indicated that MEG3, a long non-coding RNA downregulated in OS, controlled the expression of several tumor suppressor genes and oncogenes including P53, RB, MYC and TGF-β; it also negatively regulated the expression of microRNA-184 (miR-184) and down-stream effectors of the Wnt/β-catenin pathway including β-catenin, TCF4 and c-MYC [148]. Therefore, downregulation of MEG3 attenuated its tumor suppressive effect and partly resulted in the upregulation of Wnt signaling. In this model, its impact on cell metabolism relied on mTORC1-mediated activation of the S6 kinase pathway and protein synthesis. In EWS, Wnt activation was also essential for the acquisition of a metastatic phenotype and controlled a proangiogenic switch via the secretion of specific extracellular matrix (ECM) proteins called angiomatrix in a TGF-β-dependent context [47]. However, overexpression of sFRP2, a secreted Wnt antagonist, promoted osteosarcoma invasion and metastatic potential [149]. Therefore, Wnt participates at various stages of STS progression.

Loss of tumor suppressors affects several metabolic pathways

Tumor suppressors interrupt cell cycle and growth in a stressed environment, in part by regulating access to trophic pathways. Patients and mice carrying a hereditary defect in p53 (Li Fraumeni syndrome) or Rb1 (retinoblastoma) show a predisposition to sarcomas [49, 150]. In low/medium grade STS such as well-differentiated liposarcoma (WDLPS), the initial oncogenic event is the amplification of the p53 inhibitor MDM2. In DDLPS, MDM2 amplification synergizes with alterations affecting genes that regulate growth such as CDK4 and FRS2 [4], or that are required for adipocyte differentiation such as JUN, DDIT3, PTPRQ, YAP1 or CEBPA, or with alterations of DNA methylation. More generally, in STS with complex genomes (LMS, UPS, MFS, LPS, MPNST), the accumulation of frequent somatic copy number alterations (SCNAs) and/or focal mutations of TSGs leads to the deregulation of the PI3K/AKT/mTOR axis, mitosis and chromosomal maintenance [48]. As in most cancers, the timing of occurrence of p53 mutations affects tumor progression and prognosis [8, 151153]. Similarly, in mice, the combined loss of p53 [154] or CDKN2A (Ink4/Arf) [66] TSGs with oncogenic RAS lead to the development of undifferentiated STS. Conditional mutations in KRAS and p53 in muscle were sufficient to provoke high-grade STS with myofibroblastic differentiation [155].
As highlighted in Fig. 2A, p53 and AKT exert a negative feedback loop on each other, through PTEN and MDM2 regulation, respectively. Also, p53 indirectly counteracts AKT-dependent downstream effects on growth, apoptosis or metabolism [156, 157]. The tumor suppressive function of p53 depends on its role as transcription factor inducing cell cycle arrest or apoptosis via the CDKN1A (p21), or PUMA and NOXA effectors, respectively (Fig. 2A). However, in their absence, tumor suppression persists suggesting that additional mechanisms [158], including those with an impact on metabolism [156, 159] are also important (Fig. 3). Indeed, GLUT gene transcription is enhanced in STS-bearing p53 mutations [160, 161]. p53 is anti-glycolytic partly through the induction of the expression of TIGAR and PARK2 regulators [162, 163]. TIGAR dephosphorylates fructose biphosphate (FBP) into fructose-6-phosphate (F6P), shifting glucose-6-phosphate (G6P) back toward the pentose phosphate pathway (PPP). Furthermore, the cytosolic form of p53 interacts with and inhibits G6P-dehydrogenase (G6PDH) by preventing the formation of the active dimer, therefore inhibiting PPP-dependent redox control and anabolism [164] (Fig. 3). Since TIGAR expression is not strictly p53-dependent, the resulting p53 effect may be difficult to predict with regards the engagement of the PPP, but it globally interferes with glycolysis. In STS, deep deletions or more frequently amplifications of TIGAR have been documented and high TIGAR expression correlates with a better outcome [50]. PARK2/Parkin is an E3 ubiquitin ligase regulating the degradation of mitochondrial proteins. It cooperates with the mitochondrial serine/threonine kinase PINK1 and contributes to mitochondrial fitness [163, 165]. p53-mediated mitochondrial homeostasis also involves the quality control of mitochondrial DNA (mtDNA) and the expression of cell death regulators [166]. Finally, p53 induces the expression of pyruvate decarboxylase (PDC) that regenerates mitochondrial oxaloacetate and reinitiates the TCA cycle, and that of isocitrate dehydrogenase 1 (IDH1) that converts cytosolic citrate into α-ketoglutarate. In a KRAS driven-PDAC model, p53-dependent accumulation of cytosolic α-ketoglutarate activates aKGDD enzymes that regulate 5-hydroxymethylcytosine-producing TET enzymes, allowing tumor cell differentiation and growth control [167] (Fig. 3).

Metabolic fluxes and mitochondrial fitness in STS

Metabolite fluxes between organelles regulate the efficiency of various metabolic pathways in cells, but also contribute to the plasticity of metabolic adaptation.

Metabolic imbalances in STS

Through their evolution, most tumors tend to acquire metabolic features including an increase in nucleotide synthesis [168]. Investigations using PET-FDG uptake in STS patients confirmed the strong glycolytic bias documented in metastasized and poor prognosis STS [169, 170] such as ARMS [171] or ES [172]. However, these studies also revealed the considerable heterogeneity within a given tumor and between different tumor types, suggesting that the Warburg phenotype might be unstable and amenable to pharmacologic control [173]. Whereas the level of oxidative phosphorylation (OXPHOS) varies between tumors (Fig. 1), there is a general correlation between reduced mitochondrial activity, an epithelial-to-mesenchymal transition (EMT) gene signature and a poor prognosis [168]. Some STS tend to exhibit high levels of mitochondrial respiration compared to carcinomas (Fig. 1) [55, 174]. In vitro analysis of OS and RMS cell lines showed differences in the reliance on glycolysis versus respiration of tumors, with ARMS being in general less oxidative than OS or ERMS [175]. The equilibrium between glycolysis and mitochondrial respiration can be affected by various oncogenic alterations and/or metabolic requirements. Accordingly, the receptor tyrosine kinase Her4/ErbB4, an EGFR family member, is upregulated in several cancers including OS [176]. Exploration of xenograft models using untargeted metabolomics and 18F-FDG microPET/CT scan approaches showed that Her4 overexpression boosted glycolysis, glutaminolysis and OXPHOS in tumors. This hypermetabolic phenotype contributed to sustained growth and ATP production while conferring chemoresistance, as also shown in PDAC [177].
The crosstalk between metabolic pathways can also be altered in cancer as discussed in the following examples. Firstly, the upstream reaction committing glucose to glycolysis is catalyzed by phosphofructokinase-1 (PFK-1), itself allosterically inhibited by high ATP levels [178] (Fig. 3). Cancer cells express various PFK isoenzymes [179] such as the bifunctional 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatase (PFKFB) that produces F2,6-BP, thereby overriding ATP-dependent inhibition of PFK1 [180, 181] (Fig. 3). Reciprocally, an activation of a F1,6-biphosphatase such as FBP1 enhances the gluconeogenic flow and restrains glycolysis [174]. The FBP2 isoform is frequently lost in STS including LPS, FS, LMS and UPS and lower FBP2 mRNA levels correlated with poor survival in LPS [55]. In the latter study, increasing FBP2 expression impaired sarcoma cell growth, through glycolysis inhibition and induction of mitochondrial biogenesis. The latter effect was due to FBP2 nuclear translocation where, independently of its enzymatic activity, it interacted with and inhibited c-Myc-driven transcriptional activation of TFAM, an inducer of mitochondrial biogenesis [182].
Secondly, the maintenance of glycolytic flow requires the regeneration of NAD+ which originates from cytosolic lactate dehydrogenase (LDH) activity and from the malate aspartate shuttle between mitochondria and cytosol [183] (Fig. 2). By regulating NAD+ levels, mitochondrial activity limits glycolysis and consequently the Warburg effect [184]. This suggests that the persistence of mitochondrial activity can be beneficial to tumors. In addition, an LDH activity has been identified in the mitochondria where it catalyzes the aerobic oxidation of lactate into pyruvate. It is thought to contribute to the maintenance or enhancement of OXPHOS in glycolytic cells [185, 186]. Pyruvate oxidation in the mitochondria depends on PDH activity, itself inhibited by PDK. The inhibition of PDK by dichloroacetate (DCA) shifts metabolism from glycolysis to glucose oxidation and boosts ROS production as well as mitochondria-dependent apoptosis in tumors [187]. This effect is exploited in EWS and other tumors where DCA synergizes with apoptosis-inducing drugs such as cisplatin. Manipulating ROS levels appears to be a promising therapeutic approach [188]. Indeed, scavenging mitochondrial ROS (mtROS) induces p53, reduces the cell transforming potential of oncogenic RAS and in some fibrosarcoma (FS) and RMS model cell lines suppresses tumor growth [17, 189].
Thirdly, the level of mitochondrial activity depends on the availability of coenzyme A (CoA) and the acetylated form, AcCoA. CoA synthesis requires the intracellular phosphorylation of pantothenate (or vitamin B5) by pantothenate kinases [190]. Reciprocally, pantothenate derives from the recycling of food-derived or cellular CoA through an extracellular degradative process involving the vanin pantetheinases [191, 192]. Interestingly, a high vanin1 (VNN1) level correlates with a better prognosis in STS patients [66]. Lack of Vnn1 in CDKN2A deficient mice enhanced the proportion of fibrosarcomas compared to that of other cancers. In RAS-driven mouse STS lines, Vnn1 exerted an anti-Warburg effect by enhancing CoA levels and mitochondrial activity to the detriment of glycolysis, and by maintaining cell differentiation.

Mitochondrial abnormalities disrupt the TCA cycle

Mitochondrial biogenesis depends on the transcriptional coactivator PGC1α [193]. This process regulates the transition from myoblast growth to differentiation and requires a switch from the classical to the alternative NF-kB activation pathway. The latter controls PGC1α transcription [194], in cooperation with MyoD [195]. In RMS and OS models, an alteration in this switch leads to the induction of the pro-glycolytic HK2 isoform through the persistent activation of the classical NF-kB pathway [196] (Fig. 3). This might also contribute to the incomplete mitochondrial biogenesis observed in a rat RMS model featuring a deficiency in respiratory potential and poor mtROS control, thereby enhancing tumorigenesis [197].
Mutations in TCA enzymes SDH [198] and FH [95], found in STS [57], are frequent in wild-type GIST without KIT or PDGFRA mutations [152]. They provoke an interruption of the TCA cycle, uncoupled from ATP production. Consequently, excess succinate diffuses in the cytoplasm where it inhibits aKGDD enzymes involved in the regulation of epigenetic modifications, DNA repair [199] or HIF degradation, rewiring cells toward glycolysis [200]. In a mouse ovarian cancer model, targeted knockdown of Sdhb resulted in enhanced proliferation and lead to a hypermethylated epigenome promoting EMT [198]. Using metabolic tracing and SeaHorse analysis, the authors documented an increased reliance on glutamine for cell survival and a reduced mitochondrial reserve capacity, rendering cells highly sensitive to the complex I inhibitor metformin.
Mutations in IDH1 and 2 lead to the production of 2-hydroxyglutarate [201], an inducer of HIF1α stabilization. HIF1a expression and hypoxia are associated with poor survival of sarcoma patients [6870, 202204]. Hypoxia regulates apoptosis resistance, cancer stemness, metastatic properties in RMS [71, 205] and is involved in ES, GIST and LPS progression [69, 70, 203, 204]. Uncoupling of electron transport chain (ETC) complexes from ATP production does not impede anaplerotic mitochondrial uptake of glutamine, transformed into glutamate via the activity of GLS to feed the reverse metabolic flow toward citrate production and anabolism [206] (Fig. 3). Accordingly, the growth of STS subtypes overexpressing GLS is sensitive to glutamine depletion in vitro and glutaminase inhibition in vivo [59].
Some components of the ETC are encoded by mtDNA. Therefore, alterations in mtDNA may lead to respiratory defects. In OS, insufficient or altered mtDNA is associated with stressed mitochondria and enhanced tumor invasiveness [207]. In an OS cell line, ethidium bromide induced-mtDNA depletion provoked a deficiency in cytochrome oxidase and OXPHOS, leading to enhanced glycolysis and EMT [208, 209]. Furthermore, mitochondrial dysfunction and loss of transmembrane potential provoked high cytosolic Ca2+ levels, triggering calcineurin-dependent mitochondria-to-nucleus retrograde signaling that resulted in AKT activation and glycolysis [210]. In a tunable model of mitochondrial dysfunction using cytoplasmic hybrids [57], impairment of respiration lead to NADH accumulation and cytosolic recycling into NAD+ by the malate deshydrogenase pathway. NAD+ boosted glycolysis and ATP-dependent cell migration. This suggests that glycolysis-derived ATP might be preferentially used during cell migration [211]. In conclusion, there is a high intra- and inter-tumor heterogeneity in mitochondrial activity, which can be enhanced or lost, depending on the tumor context.

Tumor metabolome impacts STS progression

Metastasis

Aerobic glycolysis induced by oncogenic or hypoxic signaling provokes changes in the tumor metabolome. Lactate excretion, hypoxia-associated hypercapnia and acidification of the extracellular milieu accelerate the degradation of the extracellular matrix and facilitate metastasis [212, 213]. Reciprocally, as shown in STS [214, 215], cancer-associated fibroblasts can produce lactate and 3-hydroxybutyrate that boost cell growth, metastasis and angiogenesis when administered to tumor-bearing mice [216218]. Lactate uptake by tumors feeds their oxidative metabolism [212, 216, 219] and requires the importer MCT1, a marker of mitochondrial activity and stemness in cancer and a target gene of the fusion protein ASPSCR1/TFE3 in alveolar soft part sarcoma (ASPS) [220]. The persistence of mitochondrial activity can enhance metabolic plasticity [221], mtROS-driven anoikis, metastasis [222, 223] or resistance to therapy as shown for LPS [224]. Metabolic plasticity, required during EMT transition [225], is still incompletely documented in poorly polarized and migration-prone mesenchymal tumor cells such as sarcoma cells [226]. Indeed, hybrid epithelial/mesenchymal (E/M) phenotypes or switching from E- to N-cadherin and vimentin expression contribute to aggressiveness, metastatic properties and drug resistance [226229]. In addition, fusion protein events and translocations, frequent in childhood STS, can regulate epithelial differentiation [230, 231]. EMT is induced by cytokines such as FGFs, PDGF, TGF-β that enhance glycolysis and TCA activity [232]. TGF-β signaling synergizes with the AKT and NF-kB pathways, both potent drivers of glycolysis [233], but also antagonizes PDK4, thereby allowing pyruvate entry into the TCA. YB-1, an enhancer of HIF1α translation, is overexpressed in high-risk human sarcomas and promotes EMT and metastasis [234]. Hypoxia regulates the expression of several intracellular collagen-modifying enzymes, particularly OGDH enzymes that hydroxylate proline and lysine residues, contributing to the quality of collagen folding and the stiffness of the tissue, and thereby affecting cell migration [235]. In a UPS model, HIF1α enhances the expression of the intracellular enzyme procollagen-lysine, 2-oxoglutarate 5-dioxygenase 2 (PLOD2). Loss or overexpression of PLOD2 abrogates or restores, respectively, the metastatic potential of HIF1α-deficient tumors and human sarcomas show elevated HIF1α and PLOD2 expression in metastatic primary lesions [236]. Finally, HIF can enhance ECM degradation through the induction of various metalloproteases such as MMP or PLAUR, facilitating invasiveness.

Immunoreactivity

Several features including the level of infiltration of cytotoxic CD8+T cells or of myeloid cells, the expression of markers of immune-stimulation or -depression and the localization of these cells within the tumor, emerge as landmarks of tumor immunogenicity [8, 237240]. STS display low mutational burden as compared to other cancer types and are generally considered to be poorly immunogenic and poorly responsive to immune checkpoint blockade [8, 241]. Synovial sarcoma, soft tissue and undifferentiated LMS are the three subtypes with the lowest CNV [8, 237] and display reduced immune infiltration, virtually devoid of lymphocytes [242245]. In contrast, STS with several SCNAs, nucleotide and chromosome instabilities, such as undifferentiated LPS, MPNST [246248], AS and GIST [243] and OS [244] can present high levels of lymphoid infiltration including CD8+ T cells. Consequently, STS display a wide range of immunophenotypes [238, 249, 250]. The metabolic rewiring imposed by tumors generates a situation of competition for essential energetic resources. This concerns glucose, vitamins and essential amino acids (serine, leucine, methionine, etc.) leading to impairment of immune cell functions and memory [19, 251, 252]. The exchange of fatty acids is required for the survival of immunosuppressive myeloid cells [253] or Tregs [254], particularly under conditions of activation of the PI3K/AKT/mTOR axis that boosts lipogenesis [255]. Other metabolites such as extracellular nucleotides released upon cell death can induce immunosuppression via various mechanisms [256258]. Altogether, metabolic disturbances imposed by tumor cells directly contributes to the reorganization of the microenvironment but an exhaustive analysis of the immune landscape in STS is still lacking. Techniques such as Met-Flow [31] or SCENITH [32] should help dissecting the metabolic status of immunocytes.

Conclusions

Unraveling the complexity of sarcoma genetics has benefited from the development and improvement of multi-omics strategies. The phenotypic and molecular description of genomic alterations can now be complemented with the identification of prognostic mechanistic signatures in patients. The heterogeneity and scarcity of STS originally limited the description of their metabolic landscapes but PET-FDG analysis has contributed to their staging, prognostication and evaluation of their response to therapy. Several cell-autonomous pathways or environmental factors influence the degree of conversion toward aerobic glycolysis, justifying the use of drugs that antagonize these processes. Nevertheless, the classic distinctions between glycolytic and oxidative tumors must be carefully reconsidered; hybrid phenotypes may confer more adaptable behaviors to STS cells. Several important metabolic pathways associated with STS progression such as those of one carbon and arginine metabolisms, or the PGI pathway might provide novel therapeutic options in combination with conventional therapies. In addition, the development of novel animal models or 2D/3D culture systems has highlighted the metabolic plasticity of these tumors that may impact their energetic resources. However, these adaptations may have a price, rendering tumor cells more sensitive to combined therapies.
Oncogenic signals can lead to the expression of isoforms of glycolytic enzymes that display new functions tilting the balance between glycolysis and mitochondrial activity. Similarly, p53 can also act as a regulator of G6PDH activity, impacting biosynthetic pathways. Some metabolites such as α-ketoglutarate have emerged as key effectors of p53 action, whereas others behave as oncometabolites leading to alterations of genome integrity, metastatic behavior and therapeutic response. The balance between glycolysis, glutaminolysis and OXPHOS depends on the respective availability of key metabolites, such as amino acids, NAD+/NADH, lactate or VitB5, that regulate STS progression or differentiation. Metabolomic studies have already shown that novel metabolite signatures will complement conventional biomarkers and help stratifying prognosis and therapeutic options. The stability of the glycolytic phenotype also depends on mitochondrial activity. Alterations of mitochondrial fitness observed in STS upon alterations of mtDNA or TCA enzymes aggravate the prognostic of tumors or can affect their chemoresistance. An unstable tumor metabolome has tumor-intrinsic or extrinsic effects causing it to be pro-metastatic or immunosuppressive. Therefore, the combination of drugs targeting different metabolic pathways should impact both tumor and immune cells in a concerted manner to reinvigorate anti-tumor immunity while tilting the balance toward cell differentiation over growth.

Acknowledgements

We wish to thank Alice Carrier, Romain Roncagalli, Rafael Arguello and Jean Erland Ricci for their critical reading of the manuscript and Jonathan Ewbank for revising English expression. Guillaume Charbonnier performed the bioinformatics analysis of the TCGA database.

Declarations

Not applicable.
Not applicable.

Competing interests

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

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Unsere Produktempfehlungen

e.Med Interdisziplinär

Kombi-Abonnement

Für Ihren Erfolg in Klinik und Praxis - Die beste Hilfe in Ihrem Arbeitsalltag

Mit e.Med Interdisziplinär erhalten Sie Zugang zu allen CME-Fortbildungen und Fachzeitschriften auf SpringerMedizin.de.

e.Med Innere Medizin

Kombi-Abonnement

Mit e.Med Innere Medizin erhalten Sie Zugang zu CME-Fortbildungen des Fachgebietes Innere Medizin, den Premium-Inhalten der internistischen Fachzeitschriften, inklusive einer gedruckten internistischen Zeitschrift Ihrer Wahl.

Literatur
1.
Zurück zum Zitat Lye KL, Nordin N, Vidyadaran S, Thilakavathy K. Mesenchymal stem cells: from stem cells to sarcomas. Cell Biol Int. 2016;40:610–8.PubMedCrossRef Lye KL, Nordin N, Vidyadaran S, Thilakavathy K. Mesenchymal stem cells: from stem cells to sarcomas. Cell Biol Int. 2016;40:610–8.PubMedCrossRef
2.
4.
Zurück zum Zitat Italiano A, Di Mauro I, Rapp J, Pierron G, Auger N, Alberti L, et al. Clinical effect of molecular methods in sarcoma diagnosis (GENSARC): a prospective, multicentre, observational study. Lancet Oncol. 2016;17:532–8.PubMedCrossRef Italiano A, Di Mauro I, Rapp J, Pierron G, Auger N, Alberti L, et al. Clinical effect of molecular methods in sarcoma diagnosis (GENSARC): a prospective, multicentre, observational study. Lancet Oncol. 2016;17:532–8.PubMedCrossRef
5.
Zurück zum Zitat Skubitz KM, Skubitz AP, Xu WW, Luo X, Lagarde P, Coindre J-M, et al. Gene expression identifies heterogeneity of metastatic behavior among high-grade non-translocation associated soft tissue sarcomas. J Transl Med. 2014;12:176.PubMedPubMedCentralCrossRef Skubitz KM, Skubitz AP, Xu WW, Luo X, Lagarde P, Coindre J-M, et al. Gene expression identifies heterogeneity of metastatic behavior among high-grade non-translocation associated soft tissue sarcomas. J Transl Med. 2014;12:176.PubMedPubMedCentralCrossRef
6.
Zurück zum Zitat Grimer R, Judson I, Peake D, Seddon B. Guidelines for the management of soft tissue sarcomas. Sarcoma. 2010;2010:1–15. Grimer R, Judson I, Peake D, Seddon B. Guidelines for the management of soft tissue sarcomas. Sarcoma. 2010;2010:1–15.
7.
Zurück zum Zitat Thanindratarn P, Dean DC, Nelson SD, Hornicek FJ, Duan Z. Advances in immune checkpoint inhibitors for bone sarcoma therapy. J Bone Oncol. 2019;15:100221.PubMedPubMedCentralCrossRef Thanindratarn P, Dean DC, Nelson SD, Hornicek FJ, Duan Z. Advances in immune checkpoint inhibitors for bone sarcoma therapy. J Bone Oncol. 2019;15:100221.PubMedPubMedCentralCrossRef
8.
Zurück zum Zitat Abeshouse A, Adebamowo C, Adebamowo SN, Akbani R, Akeredolu T, Ally A, et al. Comprehensive and integrated genomic characterization of adult soft tissue sarcomas. Cell. 2017;171:950–965.e28.PubMedCentralCrossRef Abeshouse A, Adebamowo C, Adebamowo SN, Akbani R, Akeredolu T, Ally A, et al. Comprehensive and integrated genomic characterization of adult soft tissue sarcomas. Cell. 2017;171:950–965.e28.PubMedCentralCrossRef
9.
Zurück zum Zitat Guillou L, Aurias A. Soft tissue sarcomas with complex genomic profiles. Virchows Arch. 2010;456:201–17.PubMedCrossRef Guillou L, Aurias A. Soft tissue sarcomas with complex genomic profiles. Virchows Arch. 2010;456:201–17.PubMedCrossRef
10.
Zurück zum Zitat Penel N, Coindre J-M, Giraud A, Terrier P, Ranchere-Vince D, Collin F, et al. Presentation and outcome of frequent and rare sarcoma histologic subtypes: a study of 10,262 patients with localized visceral/soft tissue sarcoma managed in reference centers. Cancer. 2018;124:1179–87.PubMedCrossRef Penel N, Coindre J-M, Giraud A, Terrier P, Ranchere-Vince D, Collin F, et al. Presentation and outcome of frequent and rare sarcoma histologic subtypes: a study of 10,262 patients with localized visceral/soft tissue sarcoma managed in reference centers. Cancer. 2018;124:1179–87.PubMedCrossRef
11.
Zurück zum Zitat Taylor BS, Barretina J, Maki RG, Antonescu CR, Singer S, Ladanyi M. Advances in sarcoma genomics and new therapeutic targets. Nat Rev Cancer. 2011;11:541–57.PubMedPubMedCentralCrossRef Taylor BS, Barretina J, Maki RG, Antonescu CR, Singer S, Ladanyi M. Advances in sarcoma genomics and new therapeutic targets. Nat Rev Cancer. 2011;11:541–57.PubMedPubMedCentralCrossRef
12.
13.
Zurück zum Zitat Koppenol WH, Bounds PL, Dang CV. Otto Warburg’s contributions to current concepts of cancer metabolism. Nat Rev Cancer. 2011;11:325–37.PubMedCrossRef Koppenol WH, Bounds PL, Dang CV. Otto Warburg’s contributions to current concepts of cancer metabolism. Nat Rev Cancer. 2011;11:325–37.PubMedCrossRef
14.
Zurück zum Zitat Sinkala M, Mulder N, Patrick MD. Metabolic gene alterations impact the clinical aggressiveness and drug responses of 32 human cancers. Commun Biol. 2019;2:1–14.CrossRef Sinkala M, Mulder N, Patrick MD. Metabolic gene alterations impact the clinical aggressiveness and drug responses of 32 human cancers. Commun Biol. 2019;2:1–14.CrossRef
15.
Zurück zum Zitat Galluzzi L, Kepp O, Vander Heiden MG, Kroemer G. Metabolic targets for cancer therapy. Nat Rev Drug Discov. 2013;12:829–46.PubMedCrossRef Galluzzi L, Kepp O, Vander Heiden MG, Kroemer G. Metabolic targets for cancer therapy. Nat Rev Drug Discov. 2013;12:829–46.PubMedCrossRef
16.
Zurück zum Zitat Galluzzi L, Kepp O, Kroemer G. Mitochondria: master regulators of danger signalling. Nat Rev Mol Cell Biol. 2012;13:780–8.PubMedCrossRef Galluzzi L, Kepp O, Kroemer G. Mitochondria: master regulators of danger signalling. Nat Rev Mol Cell Biol. 2012;13:780–8.PubMedCrossRef
17.
Zurück zum Zitat Ralph SJ, Rodríguez-Enríquez S, Neuzil J, Saavedra E, Moreno-Sánchez R. The causes of cancer revisited: “Mitochondrial malignancy” and ROS-induced oncogenic transformation—why mitochondria are targets for cancer therapy. Mol Asp Med. 2010;31:145–70.CrossRef Ralph SJ, Rodríguez-Enríquez S, Neuzil J, Saavedra E, Moreno-Sánchez R. The causes of cancer revisited: “Mitochondrial malignancy” and ROS-induced oncogenic transformation—why mitochondria are targets for cancer therapy. Mol Asp Med. 2010;31:145–70.CrossRef
18.
Zurück zum Zitat Sciacovelli M, Schmidt C, Maher ER, Frezza C. Metabolic drivers in hereditary cancer syndromes. Annu Rev Cancer Biol. 2020;4:77–97.CrossRef Sciacovelli M, Schmidt C, Maher ER, Frezza C. Metabolic drivers in hereditary cancer syndromes. Annu Rev Cancer Biol. 2020;4:77–97.CrossRef
19.
Zurück zum Zitat Klein Geltink RI, Kyle RL, Pearce EL. Unraveling the complex interplay between T cell metabolism and function. Annu Rev Immunol. 2018;36:461–88.PubMedCentralCrossRef Klein Geltink RI, Kyle RL, Pearce EL. Unraveling the complex interplay between T cell metabolism and function. Annu Rev Immunol. 2018;36:461–88.PubMedCentralCrossRef
20.
Zurück zum Zitat Rosario SR, Long MD, Affronti HC, Rowsam AM, Eng KH, Smiraglia DJ. Pan-cancer analysis of transcriptional metabolic dysregulation using The Cancer Genome Atlas. Nat Commun. 2018;9:5330.PubMedPubMedCentralCrossRef Rosario SR, Long MD, Affronti HC, Rowsam AM, Eng KH, Smiraglia DJ. Pan-cancer analysis of transcriptional metabolic dysregulation using The Cancer Genome Atlas. Nat Commun. 2018;9:5330.PubMedPubMedCentralCrossRef
21.
Zurück zum Zitat Liu X, Zhang A, Fang H, Li M, Song Q, Su J, et al. Serum metabolomics strategy for understanding the therapeutic effects of Yin-Chen-Hao-Tang against Yanghuang syndrome. RSC Adv. 2018;8:7403–13.CrossRefPubMedCentral Liu X, Zhang A, Fang H, Li M, Song Q, Su J, et al. Serum metabolomics strategy for understanding the therapeutic effects of Yin-Chen-Hao-Tang against Yanghuang syndrome. RSC Adv. 2018;8:7403–13.CrossRefPubMedCentral
22.
Zurück zum Zitat Colaprico A, Silva TC, Olsen C, Garofano L, Cava C, Garolini D, et al. TCGAbiolinks: an R/Bioconductor package for integrative analysis of TCGA data. Nucleic Acids Res. 2016;44:e71–e71.PubMedCrossRef Colaprico A, Silva TC, Olsen C, Garofano L, Cava C, Garolini D, et al. TCGAbiolinks: an R/Bioconductor package for integrative analysis of TCGA data. Nucleic Acids Res. 2016;44:e71–e71.PubMedCrossRef
24.
Zurück zum Zitat Jassal B, Matthews L, Viteri G, Gong C, Lorente P, Fabregat A, et al. The reactome pathway knowledgebase. Nucleic Acids Res. 2020;48:D498-503.PubMed Jassal B, Matthews L, Viteri G, Gong C, Lorente P, Fabregat A, et al. The reactome pathway knowledgebase. Nucleic Acids Res. 2020;48:D498-503.PubMed
25.
Zurück zum Zitat Yu G, Wang L-G, Han Y, He Q-Y. clusterProfiler: an R package for comparing biological themes among gene clusters. OMICS J Integr Biol. 2012;16:284–7.CrossRef Yu G, Wang L-G, Han Y, He Q-Y. clusterProfiler: an R package for comparing biological themes among gene clusters. OMICS J Integr Biol. 2012;16:284–7.CrossRef
26.
Zurück zum Zitat Chen J, Sun M, Hua Y, Cai Z. Prognostic significance of serum lactate dehydrogenase level in osteosarcoma: a meta-analysis. J Cancer Res Clin Oncol. 2014;140:1205–10.PubMedCrossRef Chen J, Sun M, Hua Y, Cai Z. Prognostic significance of serum lactate dehydrogenase level in osteosarcoma: a meta-analysis. J Cancer Res Clin Oncol. 2014;140:1205–10.PubMedCrossRef
27.
Zurück zum Zitat Zhong Z, Mao S, Lin H, Li H, Lin J, Lin J-M. Alteration of intracellular metabolome in osteosarcoma stem cells revealed by liquid chromatography–tandem mass spectrometry. Talanta. 2019;204:6–12.PubMedCrossRef Zhong Z, Mao S, Lin H, Li H, Lin J, Lin J-M. Alteration of intracellular metabolome in osteosarcoma stem cells revealed by liquid chromatography–tandem mass spectrometry. Talanta. 2019;204:6–12.PubMedCrossRef
28.
Zurück zum Zitat Lou S, Balluff B, de Graaff MA, Cleven AHG, Bruijn IB, Bovée JVMG, et al. High-grade sarcoma diagnosis and prognosis: biomarker discovery by mass spectrometry imaging. Proteomics. 2016;16:1802–13.PubMedCrossRef Lou S, Balluff B, de Graaff MA, Cleven AHG, Bruijn IB, Bovée JVMG, et al. High-grade sarcoma diagnosis and prognosis: biomarker discovery by mass spectrometry imaging. Proteomics. 2016;16:1802–13.PubMedCrossRef
29.
Zurück zum Zitat Takahashi A, Nakayama R, Ishibashi N, Doi A, Ichinohe R, Ikuyo Y, et al. Analysis of gene expression profiles of soft tissue sarcoma using a combination of knowledge-based filtering with integration of multiple statistics. PLoS ONE. 2014;9:e106801.PubMedPubMedCentralCrossRef Takahashi A, Nakayama R, Ishibashi N, Doi A, Ichinohe R, Ikuyo Y, et al. Analysis of gene expression profiles of soft tissue sarcoma using a combination of knowledge-based filtering with integration of multiple statistics. PLoS ONE. 2014;9:e106801.PubMedPubMedCentralCrossRef
30.
Zurück zum Zitat Sun C, Li T, Song X, Huang L, Zang Q, Xu J, et al. Spatially resolved metabolomics to discover tumor-associated metabolic alterations. PNAS. 2019;116:52–7.PubMedCrossRef Sun C, Li T, Song X, Huang L, Zang Q, Xu J, et al. Spatially resolved metabolomics to discover tumor-associated metabolic alterations. PNAS. 2019;116:52–7.PubMedCrossRef
31.
Zurück zum Zitat Ahl PJ, Hopkins RA, Xiang WW, Au B, Kaliaperumal N, Fairhurst A-M, et al. Met-Flow, a strategy for single-cell metabolic analysis highlights dynamic changes in immune subpopulations. Commun Biol. 2020;3:1–15.CrossRef Ahl PJ, Hopkins RA, Xiang WW, Au B, Kaliaperumal N, Fairhurst A-M, et al. Met-Flow, a strategy for single-cell metabolic analysis highlights dynamic changes in immune subpopulations. Commun Biol. 2020;3:1–15.CrossRef
32.
Zurück zum Zitat Argüello RJ, Combes AJ, Char R, Gigan J-P, Baaziz AI, Bousiquot E, et al. SCENITH: a flow cytometry-based method to functionally profile energy metabolism with single-cell resolution. Cell Metab. 2020;32:1063-1075.e7.PubMedCrossRefPubMedCentral Argüello RJ, Combes AJ, Char R, Gigan J-P, Baaziz AI, Bousiquot E, et al. SCENITH: a flow cytometry-based method to functionally profile energy metabolism with single-cell resolution. Cell Metab. 2020;32:1063-1075.e7.PubMedCrossRefPubMedCentral
33.
Zurück zum Zitat Serrano C, Romagosa C, Hernández-Losa J, Simonetti S, Valverde C, Moliné T, et al. RAS/MAPK pathway hyperactivation determines poor prognosis in undifferentiated pleomorphic sarcomas. Cancer. 2016;122:99–107.PubMedCrossRef Serrano C, Romagosa C, Hernández-Losa J, Simonetti S, Valverde C, Moliné T, et al. RAS/MAPK pathway hyperactivation determines poor prognosis in undifferentiated pleomorphic sarcomas. Cancer. 2016;122:99–107.PubMedCrossRef
34.
Zurück zum Zitat Dodd RD. Emerging targets in sarcoma: rising to the challenge of RAS signaling in undifferentiated pleomorphic sarcoma: RAS/MAPK levels in UPS. Cancer. 2016;122:17–9.PubMedCrossRef Dodd RD. Emerging targets in sarcoma: rising to the challenge of RAS signaling in undifferentiated pleomorphic sarcoma: RAS/MAPK levels in UPS. Cancer. 2016;122:17–9.PubMedCrossRef
35.
Zurück zum Zitat Mora J, Rodríguez E, de Torres C, Cardesa T, Ríos J, Hernández T, et al. Activated growth signaling pathway expression in Ewing sarcoma and clinical outcome. Pediatr Blood Cancer. 2012;58:532–8.PubMedCrossRef Mora J, Rodríguez E, de Torres C, Cardesa T, Ríos J, Hernández T, et al. Activated growth signaling pathway expression in Ewing sarcoma and clinical outcome. Pediatr Blood Cancer. 2012;58:532–8.PubMedCrossRef
36.
Zurück zum Zitat Ahmed AA, Sherman AK, Pawel BR. Expression of therapeutic targets in Ewing sarcoma family tumors. Hum Pathol. 2012;43:1077–83.PubMedCrossRef Ahmed AA, Sherman AK, Pawel BR. Expression of therapeutic targets in Ewing sarcoma family tumors. Hum Pathol. 2012;43:1077–83.PubMedCrossRef
37.
Zurück zum Zitat Noh B-J, Jung W-W, Kim H-S, Park Y-K. Pathogenetic implications of early growth response 1 in Ewing sarcoma. Pathology. 2019;51:605–9.PubMedCrossRef Noh B-J, Jung W-W, Kim H-S, Park Y-K. Pathogenetic implications of early growth response 1 in Ewing sarcoma. Pathology. 2019;51:605–9.PubMedCrossRef
38.
Zurück zum Zitat Machado I, López-Guerrero JA, Scotlandi K, Picci P, Llombart-Bosch A. Immunohistochemical analysis and prognostic significance of PD-L1, PD-1, and CD8+ tumor-infiltrating lymphocytes in Ewing’s sarcoma family of tumors (ESFT). Virchows Arch. 2018;472:815–24.PubMedCrossRef Machado I, López-Guerrero JA, Scotlandi K, Picci P, Llombart-Bosch A. Immunohistochemical analysis and prognostic significance of PD-L1, PD-1, and CD8+ tumor-infiltrating lymphocytes in Ewing’s sarcoma family of tumors (ESFT). Virchows Arch. 2018;472:815–24.PubMedCrossRef
39.
Zurück zum Zitat Glorie N, Baert T. Circulating Protein biomarkers to differentiate uterine sarcomas from leiomyomas. Anticancer Res. 2019;39:3981–9.PubMedCrossRef Glorie N, Baert T. Circulating Protein biomarkers to differentiate uterine sarcomas from leiomyomas. Anticancer Res. 2019;39:3981–9.PubMedCrossRef
41.
Zurück zum Zitat Mohamed AD, Tremblay AM, Murray GI, Wackerhage H. The Hippo signal transduction pathway in soft tissue sarcomas. Biochim Biophys Acta (BBA) Rev Cancer. 2015;1856:121–9.CrossRef Mohamed AD, Tremblay AM, Murray GI, Wackerhage H. The Hippo signal transduction pathway in soft tissue sarcomas. Biochim Biophys Acta (BBA) Rev Cancer. 2015;1856:121–9.CrossRef
42.
Zurück zum Zitat Crose LES, Galindo KA, Kephart JG, Chen C, Fitamant J, Bardeesy N, et al. Alveolar rhabdomyosarcoma-associated PAX3-FOXO1 promotes tumorigenesis via Hippo pathway suppression. J Clin Investig. 2014;124:285–96.PubMedCrossRef Crose LES, Galindo KA, Kephart JG, Chen C, Fitamant J, Bardeesy N, et al. Alveolar rhabdomyosarcoma-associated PAX3-FOXO1 promotes tumorigenesis via Hippo pathway suppression. J Clin Investig. 2014;124:285–96.PubMedCrossRef
43.
Zurück zum Zitat Fullenkamp CA, Hall SL, Jaber OI, Pakalniskis BL, Savage EC, Savage JM, et al. TAZ and YAP are frequently activated oncoproteins in sarcomas. Oncotarget. 2016;7:30094–108.PubMedPubMedCentralCrossRef Fullenkamp CA, Hall SL, Jaber OI, Pakalniskis BL, Savage EC, Savage JM, et al. TAZ and YAP are frequently activated oncoproteins in sarcomas. Oncotarget. 2016;7:30094–108.PubMedPubMedCentralCrossRef
44.
Zurück zum Zitat Trautmann M, Cheng Y, Jensen P, Azoitei N, Brunner I, Hüllein J, et al. Requirement for YAP1 signaling in myxoid liposarcoma. EMBO Mol Med. 2019;11:e9889.PubMedPubMedCentralCrossRef Trautmann M, Cheng Y, Jensen P, Azoitei N, Brunner I, Hüllein J, et al. Requirement for YAP1 signaling in myxoid liposarcoma. EMBO Mol Med. 2019;11:e9889.PubMedPubMedCentralCrossRef
45.
Zurück zum Zitat Tremblay AM, Missiaglia E, Galli GG, Hettmer S, Urcia R, Carrara M, et al. The Hippo transducer YAP1 transforms activated satellite cells and is a potent effector of embryonal rhabdomyosarcoma formation. Cancer Cell. 2014;26:273–87.PubMedCrossRef Tremblay AM, Missiaglia E, Galli GG, Hettmer S, Urcia R, Carrara M, et al. The Hippo transducer YAP1 transforms activated satellite cells and is a potent effector of embryonal rhabdomyosarcoma formation. Cancer Cell. 2014;26:273–87.PubMedCrossRef
46.
Zurück zum Zitat Pedersen EA, Menon R, Bailey KM, Thomas DG, Van Noord RA, Tran J, et al. Activation of Wnt/-catenin in Ewing sarcoma cells antagonizes EWS/ETS function and promotes phenotypic transition to more metastatic cell states. Can Res. 2016;76:5040–53.CrossRef Pedersen EA, Menon R, Bailey KM, Thomas DG, Van Noord RA, Tran J, et al. Activation of Wnt/-catenin in Ewing sarcoma cells antagonizes EWS/ETS function and promotes phenotypic transition to more metastatic cell states. Can Res. 2016;76:5040–53.CrossRef
47.
Zurück zum Zitat Hawkins AG, Pedersen EA, Treichel S, Temprine K, Sperring C, Read JA, et al. Wnt/β-catenin-activated Ewing sarcoma cells promote the angiogenic switch. JCI Insight. 2020;5:e135188.PubMedCentralCrossRef Hawkins AG, Pedersen EA, Treichel S, Temprine K, Sperring C, Read JA, et al. Wnt/β-catenin-activated Ewing sarcoma cells promote the angiogenic switch. JCI Insight. 2020;5:e135188.PubMedCentralCrossRef
48.
Zurück zum Zitat Chibon F, Lagarde P, Salas S, Pérot G, Brouste V, Tirode F, et al. Validated prediction of clinical outcome in sarcomas and multiple types of cancer on the basis of a gene expression signature related to genome complexity. Nat Med. 2010;16:781–7.PubMedCrossRef Chibon F, Lagarde P, Salas S, Pérot G, Brouste V, Tirode F, et al. Validated prediction of clinical outcome in sarcomas and multiple types of cancer on the basis of a gene expression signature related to genome complexity. Nat Med. 2010;16:781–7.PubMedCrossRef
49.
Zurück zum Zitat Ballinger ML, Goode DL, Ray-Coquard I, James PA, Mitchell G, Niedermayr E, et al. Monogenic and polygenic determinants of sarcoma risk: an international genetic study. Lancet Oncol. 2016;17:1261–71.PubMedCrossRef Ballinger ML, Goode DL, Ray-Coquard I, James PA, Mitchell G, Niedermayr E, et al. Monogenic and polygenic determinants of sarcoma risk: an international genetic study. Lancet Oncol. 2016;17:1261–71.PubMedCrossRef
50.
Zurück zum Zitat Fang P, de Souza C, Minn K, Chien J. Genome-scale CRISPR knockout screen identifies TIGAR as a modifier of PARP inhibitor sensitivity. Commun Biol. 2019;2:16.CrossRef Fang P, de Souza C, Minn K, Chien J. Genome-scale CRISPR knockout screen identifies TIGAR as a modifier of PARP inhibitor sensitivity. Commun Biol. 2019;2:16.CrossRef
51.
52.
Zurück zum Zitat de Necochea-Campion R, Zuckerman LM, Mirshahidi HR, Khosrowpour S, Chen C-S, Mirshahidi S. Metastatic biomarkers in synovial sarcoma. Biomark Res. 2017;5:1–8.CrossRef de Necochea-Campion R, Zuckerman LM, Mirshahidi HR, Khosrowpour S, Chen C-S, Mirshahidi S. Metastatic biomarkers in synovial sarcoma. Biomark Res. 2017;5:1–8.CrossRef
53.
Zurück zum Zitat Benz MR, Dry SM, Eilber FC, Allen-Auerbach MS, Tap WD, Elashoff D, et al. Correlation between glycolytic phenotype and tumor grade in soft-tissue sarcomas by 18F-FDG PET. J Nucl Med. 2010;51:1174–81.PubMedCrossRef Benz MR, Dry SM, Eilber FC, Allen-Auerbach MS, Tap WD, Elashoff D, et al. Correlation between glycolytic phenotype and tumor grade in soft-tissue sarcomas by 18F-FDG PET. J Nucl Med. 2010;51:1174–81.PubMedCrossRef
54.
Zurück zum Zitat Chen H, Chen Y, Liu H, Que Y, Zhang X, Zheng F. Integrated expression profiles analysis reveals correlations between the IL-33/ST2 axis and CD8+ T cells, regulatory T cells, and myeloid-derived suppressor cells in soft tissue sarcoma. Front Immunol. 2018;9:1179.PubMedPubMedCentralCrossRef Chen H, Chen Y, Liu H, Que Y, Zhang X, Zheng F. Integrated expression profiles analysis reveals correlations between the IL-33/ST2 axis and CD8+ T cells, regulatory T cells, and myeloid-derived suppressor cells in soft tissue sarcoma. Front Immunol. 2018;9:1179.PubMedPubMedCentralCrossRef
55.
Zurück zum Zitat Huangyang P, Li F, Lee P, Nissim I, Weljie AM, Mancuso A, et al. Fructose-1,6-bisphosphatase 2 inhibits sarcoma progression by restraining mitochondrial biogenesis. Cell Metab. 2020;31:174-188.e7.PubMedCrossRef Huangyang P, Li F, Lee P, Nissim I, Weljie AM, Mancuso A, et al. Fructose-1,6-bisphosphatase 2 inhibits sarcoma progression by restraining mitochondrial biogenesis. Cell Metab. 2020;31:174-188.e7.PubMedCrossRef
56.
Zurück zum Zitat Pillozzi S, Bernini A, Palchetti I, Crociani O, Antonuzzo L, Campanacci D, et al. Soft tissue sarcoma: an insight on biomarkers at molecular metabolic and cellular level. Cancers. 2021;13:3044.PubMedPubMedCentralCrossRef Pillozzi S, Bernini A, Palchetti I, Crociani O, Antonuzzo L, Campanacci D, et al. Soft tissue sarcoma: an insight on biomarkers at molecular metabolic and cellular level. Cancers. 2021;13:3044.PubMedPubMedCentralCrossRef
57.
Zurück zum Zitat Gaude E, Schmidt C, Gammage PA, Dugourd A, Blacker T, Chew SP, et al. NADH shuttling couples cytosolic reductive carboxylation of glutamine with glycolysis in cells with mitochondrial dysfunction. Mol Cell. 2018;69:581-593.e7.PubMedPubMedCentralCrossRef Gaude E, Schmidt C, Gammage PA, Dugourd A, Blacker T, Chew SP, et al. NADH shuttling couples cytosolic reductive carboxylation of glutamine with glycolysis in cells with mitochondrial dysfunction. Mol Cell. 2018;69:581-593.e7.PubMedPubMedCentralCrossRef
58.
Zurück zum Zitat Zhang B, Yang L, Wang X, Fu D. Identification of a survival-related signature for sarcoma patients through integrated transcriptomic and proteomic profiling analyses. Gene. 2021;764:145105.PubMedCrossRef Zhang B, Yang L, Wang X, Fu D. Identification of a survival-related signature for sarcoma patients through integrated transcriptomic and proteomic profiling analyses. Gene. 2021;764:145105.PubMedCrossRef
59.
Zurück zum Zitat Lee P, Malik D, Perkons N, Huangyang P, Khare S, Rhoades S, et al. Targeting glutamine metabolism slows soft tissue sarcoma growth. Nat Commun. 2020;11:498.PubMedPubMedCentralCrossRef Lee P, Malik D, Perkons N, Huangyang P, Khare S, Rhoades S, et al. Targeting glutamine metabolism slows soft tissue sarcoma growth. Nat Commun. 2020;11:498.PubMedPubMedCentralCrossRef
60.
Zurück zum Zitat Sen N, Cross AM, Lorenzi PL, Khan J, Gryder BE, Kim S, et al. EWS-FLI1 reprograms the metabolism of Ewing sarcoma cells via positive regulation of glutamine import and serine-glycine biosynthesis. Mol Carcinog. 2018;57:1342–57.PubMedPubMedCentralCrossRef Sen N, Cross AM, Lorenzi PL, Khan J, Gryder BE, Kim S, et al. EWS-FLI1 reprograms the metabolism of Ewing sarcoma cells via positive regulation of glutamine import and serine-glycine biosynthesis. Mol Carcinog. 2018;57:1342–57.PubMedPubMedCentralCrossRef
61.
Zurück zum Zitat Huang H-Y, Wu W-R, Wang Y-H, Wang J-W, Fang F-M, Tsai J-W, et al. ASS1 as a novel tumor suppressor gene in myxofibrosarcoma: aberrant loss via epigenetic DNA methylation confers aggressive phenotypes, negative prognostic impact, and therapeutic relevance. Clin Cancer Res. 2013;19:2861–72.PubMedCrossRef Huang H-Y, Wu W-R, Wang Y-H, Wang J-W, Fang F-M, Tsai J-W, et al. ASS1 as a novel tumor suppressor gene in myxofibrosarcoma: aberrant loss via epigenetic DNA methylation confers aggressive phenotypes, negative prognostic impact, and therapeutic relevance. Clin Cancer Res. 2013;19:2861–72.PubMedCrossRef
62.
Zurück zum Zitat Li T, Zhu Y, Cheng F, Lu C, Jung JU, Gao S-J. Oncogenic Kaposi’s sarcoma-associated herpesvirus upregulates argininosuccinate synthase 1, a rate-limiting enzyme of the citrulline-nitric oxide cycle, to activate the STAT3 pathway and promote growth transformation. J Virol. 2019;93:e01599-e1618.PubMedPubMedCentral Li T, Zhu Y, Cheng F, Lu C, Jung JU, Gao S-J. Oncogenic Kaposi’s sarcoma-associated herpesvirus upregulates argininosuccinate synthase 1, a rate-limiting enzyme of the citrulline-nitric oxide cycle, to activate the STAT3 pathway and promote growth transformation. J Virol. 2019;93:e01599-e1618.PubMedPubMedCentral
63.
Zurück zum Zitat Choi YM, Yeo HK, Park YW, Lee JY. Structural analysis of thymidylate synthase from Kaposi’s sarcoma-associated herpesvirus with the anticancer drug raltitrexed. PLoS ONE. 2016;11:e0168019.PubMedPubMedCentralCrossRef Choi YM, Yeo HK, Park YW, Lee JY. Structural analysis of thymidylate synthase from Kaposi’s sarcoma-associated herpesvirus with the anticancer drug raltitrexed. PLoS ONE. 2016;11:e0168019.PubMedPubMedCentralCrossRef
64.
Zurück zum Zitat Tanner JM, Bensard C, Wei P, Krah NM, Schell JC, Gardiner J, et al. EWS/FLI is a master regulator of metabolic reprogramming in Ewing sarcoma. Mol Cancer Res. 2017;15:1517–30.PubMedPubMedCentralCrossRef Tanner JM, Bensard C, Wei P, Krah NM, Schell JC, Gardiner J, et al. EWS/FLI is a master regulator of metabolic reprogramming in Ewing sarcoma. Mol Cancer Res. 2017;15:1517–30.PubMedPubMedCentralCrossRef
65.
Zurück zum Zitat Mutz CN, Schwentner R, Kauer MO, Katschnig AM, Kromp F, Aryee DNT, et al. EWS-FLI1 impairs aryl hydrocarbon receptor activation by blocking tryptophan breakdown via the kynurenine pathway. FEBS Lett. 2016;590:2063–75.PubMedPubMedCentralCrossRef Mutz CN, Schwentner R, Kauer MO, Katschnig AM, Kromp F, Aryee DNT, et al. EWS-FLI1 impairs aryl hydrocarbon receptor activation by blocking tryptophan breakdown via the kynurenine pathway. FEBS Lett. 2016;590:2063–75.PubMedPubMedCentralCrossRef
66.
Zurück zum Zitat Giessner C, Millet V, Mostert KJ, Gensollen T, Vu Manh T-P, Garibal M, et al. Vnn1 pantetheinase limits the Warburg effect and sarcoma growth by rescuing mitochondrial activity. Life Sci Alliance. 2018;1:e201800073.PubMedPubMedCentralCrossRef Giessner C, Millet V, Mostert KJ, Gensollen T, Vu Manh T-P, Garibal M, et al. Vnn1 pantetheinase limits the Warburg effect and sarcoma growth by rescuing mitochondrial activity. Life Sci Alliance. 2018;1:e201800073.PubMedPubMedCentralCrossRef
67.
Zurück zum Zitat Lou S, Balluff B, Cleven AHG, Bovée JVMG, McDonnell LA. Prognostic metabolite biomarkers for soft tissue sarcomas discovered by mass spectrometry imaging. J Am Soc Mass Spectrom. 2017;28:376–83.PubMedCrossRef Lou S, Balluff B, Cleven AHG, Bovée JVMG, McDonnell LA. Prognostic metabolite biomarkers for soft tissue sarcomas discovered by mass spectrometry imaging. J Am Soc Mass Spectrom. 2017;28:376–83.PubMedCrossRef
68.
Zurück zum Zitat Li Y, Zhang W, Li S, Tu C. Prognosis value of Hypoxia-inducible factor-1α expression in patients with bone and soft tissue sarcoma: a meta-analysis. Springerplus. 2016;5:1–10. Li Y, Zhang W, Li S, Tu C. Prognosis value of Hypoxia-inducible factor-1α expression in patients with bone and soft tissue sarcoma: a meta-analysis. Springerplus. 2016;5:1–10.
69.
Zurück zum Zitat Aryee DNT, Niedan S, Kauer M, Schwentner R, Bennani-Baiti IM, Ban J, et al. Hypoxia modulates EWS-FLI1 transcriptional signature and enhances the malignant properties of Ewing’s sarcoma cells in vitro. Can Res. 2010;70:4015–23.CrossRef Aryee DNT, Niedan S, Kauer M, Schwentner R, Bennani-Baiti IM, Ban J, et al. Hypoxia modulates EWS-FLI1 transcriptional signature and enhances the malignant properties of Ewing’s sarcoma cells in vitro. Can Res. 2010;70:4015–23.CrossRef
70.
Zurück zum Zitat Jham BC, Ma T, Hu J, Chaisuparat R, Friedman ER, Pandolfi PP, et al. Amplification of the angiogenic signal through the activation of the TSC/mTOR/HIF axis by the KSHV vGPCR in Kaposi’s sarcoma. PLoS ONE. 2011;6:e19103.PubMedPubMedCentralCrossRef Jham BC, Ma T, Hu J, Chaisuparat R, Friedman ER, Pandolfi PP, et al. Amplification of the angiogenic signal through the activation of the TSC/mTOR/HIF axis by the KSHV vGPCR in Kaposi’s sarcoma. PLoS ONE. 2011;6:e19103.PubMedPubMedCentralCrossRef
71.
Zurück zum Zitat Kilic M, Kasperczyk H, Fulda S, Debatin K-M. Role of hypoxia inducible factor-1 alpha in modulation of apoptosis resistance. Oncogene. 2007;26:2027–38.PubMedCrossRef Kilic M, Kasperczyk H, Fulda S, Debatin K-M. Role of hypoxia inducible factor-1 alpha in modulation of apoptosis resistance. Oncogene. 2007;26:2027–38.PubMedCrossRef
72.
Zurück zum Zitat Merry E. Predictive and prognostic transcriptomic biomarkers in soft tissue sarcomas. npj Precis Oncol. 2021;5:1–8. Merry E. Predictive and prognostic transcriptomic biomarkers in soft tissue sarcomas. npj Precis Oncol. 2021;5:1–8.
73.
Zurück zum Zitat Zwang Y, Oren M, Yarden Y. Consistency test of the cell cycle: roles for p53 and EGR1. Can Res. 2012;72:1051–4.CrossRef Zwang Y, Oren M, Yarden Y. Consistency test of the cell cycle: roles for p53 and EGR1. Can Res. 2012;72:1051–4.CrossRef
74.
Zurück zum Zitat Jones SM, Kazlauskas A. Growth-factor-dependent mitogenesis requires two distinct phases of signalling. Nat Cell Biol. 2001;3:165–72.PubMedCrossRef Jones SM, Kazlauskas A. Growth-factor-dependent mitogenesis requires two distinct phases of signalling. Nat Cell Biol. 2001;3:165–72.PubMedCrossRef
76.
Zurück zum Zitat Janku F, Hong DS, Fu S, Piha-Paul SA, Naing A, Falchook GS, et al. Assessing PIK3CA and PTEN in early-phase trials with PI3K/AKT/mTOR inhibitors. Cell Rep. 2014;6:377–87.PubMedPubMedCentralCrossRef Janku F, Hong DS, Fu S, Piha-Paul SA, Naing A, Falchook GS, et al. Assessing PIK3CA and PTEN in early-phase trials with PI3K/AKT/mTOR inhibitors. Cell Rep. 2014;6:377–87.PubMedPubMedCentralCrossRef
77.
Zurück zum Zitat García-Valverde A, Rosell J, Serna G, Valverde C, Carles J, Nuciforo P, et al. Preclinical activity of PI3K inhibitor copanlisib in gastrointestinal stromal tumor. Mol Cancer Ther. 2020;19:1289–97.PubMedCrossRef García-Valverde A, Rosell J, Serna G, Valverde C, Carles J, Nuciforo P, et al. Preclinical activity of PI3K inhibitor copanlisib in gastrointestinal stromal tumor. Mol Cancer Ther. 2020;19:1289–97.PubMedCrossRef
78.
Zurück zum Zitat Damodaran S, Zhao F, Deming DA, Mitchell EP, Wright JJ, Doyle LA, et al. Phase II study of copanlisib in patients with tumors with PIK3CA mutations (PTEN loss allowed): NCI MATCH EAY131-Z1F. JCO Wolters Kluwer. 2020;38:3506–3506. Damodaran S, Zhao F, Deming DA, Mitchell EP, Wright JJ, Doyle LA, et al. Phase II study of copanlisib in patients with tumors with PIK3CA mutations (PTEN loss allowed): NCI MATCH EAY131-Z1F. JCO Wolters Kluwer. 2020;38:3506–3506.
79.
Zurück zum Zitat Pollack SM, Ingham M, Spraker MB, Schwartz GK. Emerging targeted and immune-based therapies in sarcoma. JCO. 2018;36:125–35.CrossRef Pollack SM, Ingham M, Spraker MB, Schwartz GK. Emerging targeted and immune-based therapies in sarcoma. JCO. 2018;36:125–35.CrossRef
80.
Zurück zum Zitat Babichev Y, Kabaroff L, Datti A, Uehling D, Isaac M, Al-awar R, et al. PI3K/AKT/mTOR inhibition in combination with doxorubicin is an effective therapy for leiomyosarcoma. J Transl Med. 2016;14:1–12.CrossRef Babichev Y, Kabaroff L, Datti A, Uehling D, Isaac M, Al-awar R, et al. PI3K/AKT/mTOR inhibition in combination with doxorubicin is an effective therapy for leiomyosarcoma. J Transl Med. 2016;14:1–12.CrossRef
81.
Zurück zum Zitat Desai C, Thomason J, Kohlmeyer JL, Reisetter AC, Ahirwar P, Jahanseir K, et al. Prognostic and therapeutic value of the Hippo pathway, RABL6A, and p53-MDM2 axes in sarcomas. Oncotarget Impact J. 2021;12:740–55.CrossRef Desai C, Thomason J, Kohlmeyer JL, Reisetter AC, Ahirwar P, Jahanseir K, et al. Prognostic and therapeutic value of the Hippo pathway, RABL6A, and p53-MDM2 axes in sarcomas. Oncotarget Impact J. 2021;12:740–55.CrossRef
82.
Zurück zum Zitat Rytlewski JD, Scalora N, Garcia K, Tanas M, Toor F, Miller B, et al. Photodynamic therapy using Hippo pathway inhibitor verteporfin: a potential dual mechanistic approach in treatment of soft tissue sarcomas. Cancers (Basel). 2021;13:675.CrossRef Rytlewski JD, Scalora N, Garcia K, Tanas M, Toor F, Miller B, et al. Photodynamic therapy using Hippo pathway inhibitor verteporfin: a potential dual mechanistic approach in treatment of soft tissue sarcomas. Cancers (Basel). 2021;13:675.CrossRef
83.
Zurück zum Zitat Isfort I, Cyra M, Elges S, Kailayangiri S, Altvater B, Rossig C, et al. SS18-SSX-dependent YAP/TAZ signaling in synovial sarcoma. Clin Cancer Res. 2019;25:3718–31.PubMedCrossRef Isfort I, Cyra M, Elges S, Kailayangiri S, Altvater B, Rossig C, et al. SS18-SSX-dependent YAP/TAZ signaling in synovial sarcoma. Clin Cancer Res. 2019;25:3718–31.PubMedCrossRef
84.
Zurück zum Zitat Bierbaumer L, Katschnig AM, Radic-Sarikas B, Kauer MO, Petro JA, Högler S, et al. YAP/TAZ inhibition reduces metastatic potential of Ewing sarcoma cells. Oncogenesis. 2021;10:1–13.CrossRef Bierbaumer L, Katschnig AM, Radic-Sarikas B, Kauer MO, Petro JA, Högler S, et al. YAP/TAZ inhibition reduces metastatic potential of Ewing sarcoma cells. Oncogenesis. 2021;10:1–13.CrossRef
85.
Zurück zum Zitat Tap WD, Villalobos VM, Cote GM, Burris H, Janku F, Mir O, et al. Phase I study of the mutant IDH1 inhibitor ivosidenib: safety and clinical activity in patients with advanced chondrosarcoma. J Clin Oncol. 2020;38:1693–701.PubMedPubMedCentralCrossRef Tap WD, Villalobos VM, Cote GM, Burris H, Janku F, Mir O, et al. Phase I study of the mutant IDH1 inhibitor ivosidenib: safety and clinical activity in patients with advanced chondrosarcoma. J Clin Oncol. 2020;38:1693–701.PubMedPubMedCentralCrossRef
86.
Zurück zum Zitat Lee D, Omofoye OA, Karnati T, Graff JP, Shahlaie K. Intracranial myeloid sarcoma presentation in distant acute myeloid leukemia remission. J Clin Neurosci. 2021;89:158–60.PubMedCrossRef Lee D, Omofoye OA, Karnati T, Graff JP, Shahlaie K. Intracranial myeloid sarcoma presentation in distant acute myeloid leukemia remission. J Clin Neurosci. 2021;89:158–60.PubMedCrossRef
87.
Zurück zum Zitat Cojocaru E, Wilding C, Engelman B, Huang P, Jones RL. Is the IDH mutation a good target for chondrosarcoma treatment? Curr Mol Biol Rep. 2020;6:1–9.CrossRef Cojocaru E, Wilding C, Engelman B, Huang P, Jones RL. Is the IDH mutation a good target for chondrosarcoma treatment? Curr Mol Biol Rep. 2020;6:1–9.CrossRef
88.
Zurück zum Zitat Bean GR, Kremer JC, Prudner BC, Schenone AD, Yao J-C, Schultze MB, et al. A metabolic synthetic lethal strategy with arginine deprivation and chloroquine leads to cell death in ASS1-deficient sarcomas. Cell Death Dis. 2016;7:e2406–e2406.PubMedPubMedCentralCrossRef Bean GR, Kremer JC, Prudner BC, Schenone AD, Yao J-C, Schultze MB, et al. A metabolic synthetic lethal strategy with arginine deprivation and chloroquine leads to cell death in ASS1-deficient sarcomas. Cell Death Dis. 2016;7:e2406–e2406.PubMedPubMedCentralCrossRef
89.
Zurück zum Zitat Kremer JC, Prudner BC, Lange SES, Bean GR, Schultze MB, Brashears CB, et al. Arginine deprivation inhibits the Warburg effect and upregulates glutamine anaplerosis and serine biosynthesis in ASS1-deficient cancers. Cell Rep. 2017;18:991–1004.PubMedPubMedCentralCrossRef Kremer JC, Prudner BC, Lange SES, Bean GR, Schultze MB, Brashears CB, et al. Arginine deprivation inhibits the Warburg effect and upregulates glutamine anaplerosis and serine biosynthesis in ASS1-deficient cancers. Cell Rep. 2017;18:991–1004.PubMedPubMedCentralCrossRef
90.
Zurück zum Zitat Rooke M, Coupland LA, Truong T, Blackburn AC. Dichloroacetate is an effective treatment for sarcoma models in vitro and in vivo. Cancer Metab. 2014;2:P9.PubMedCentralCrossRef Rooke M, Coupland LA, Truong T, Blackburn AC. Dichloroacetate is an effective treatment for sarcoma models in vitro and in vivo. Cancer Metab. 2014;2:P9.PubMedCentralCrossRef
92.
Zurück zum Zitat Travelli C, Consonni FM, Sangaletti S, Storto M, Morlacchi S, Grolla AA, et al. Nicotinamide phosphoribosyltransferase acts as a metabolic gate for mobilization of myeloid-derived suppressor cells. Cancer Res. 2019;79:1938–51.PubMedCrossRef Travelli C, Consonni FM, Sangaletti S, Storto M, Morlacchi S, Grolla AA, et al. Nicotinamide phosphoribosyltransferase acts as a metabolic gate for mobilization of myeloid-derived suppressor cells. Cancer Res. 2019;79:1938–51.PubMedCrossRef
93.
Zurück zum Zitat Hirota S. Gain-of-function mutations of c-kit in human gastrointestinal stromal tumors. Science. 1998;279:577–80.PubMedCrossRef Hirota S. Gain-of-function mutations of c-kit in human gastrointestinal stromal tumors. Science. 1998;279:577–80.PubMedCrossRef
94.
Zurück zum Zitat Goncalves MD, Hopkins BD, Cantley LC. Phosphatidylinositol 3-kinase, growth disorders, and cancer. N Engl J Med. 2018;379:2052.PubMedCrossRef Goncalves MD, Hopkins BD, Cantley LC. Phosphatidylinositol 3-kinase, growth disorders, and cancer. N Engl J Med. 2018;379:2052.PubMedCrossRef
96.
Zurück zum Zitat Lee C-L, Mowery YM, Daniel AR, Zhang D, Sibley AB, Delaney JR, et al. Mutational landscape in genetically engineered, carcinogen-induced, and radiation-induced mouse sarcoma. JCI Insight. 2019;4:e128698.PubMedCentralCrossRef Lee C-L, Mowery YM, Daniel AR, Zhang D, Sibley AB, Delaney JR, et al. Mutational landscape in genetically engineered, carcinogen-induced, and radiation-induced mouse sarcoma. JCI Insight. 2019;4:e128698.PubMedCentralCrossRef
97.
Zurück zum Zitat Chen X, Stewart E, Shelat AA, Qu C, Bahrami A, Hatley M, et al. Targeting oxidative stress in embryonal rhabdomyosarcoma. Cancer Cell. 2013;24:710–24.PubMedPubMedCentralCrossRef Chen X, Stewart E, Shelat AA, Qu C, Bahrami A, Hatley M, et al. Targeting oxidative stress in embryonal rhabdomyosarcoma. Cancer Cell. 2013;24:710–24.PubMedPubMedCentralCrossRef
98.
Zurück zum Zitat Pal A, Chiu HY, Taneja R. Genetics, epigenetics and redox homeostasis in rhabdomyosarcoma: emerging targets and therapeutics. Redox Biol. 2019;25:101124.PubMedPubMedCentralCrossRef Pal A, Chiu HY, Taneja R. Genetics, epigenetics and redox homeostasis in rhabdomyosarcoma: emerging targets and therapeutics. Redox Biol. 2019;25:101124.PubMedPubMedCentralCrossRef
100.
Zurück zum Zitat Tanner LB, Goglia AG, Wei MH, Sehgal T, Parsons LR, Park JO, et al. Four key steps control glycolytic flux in mammalian cells. Cell Syst. 2018;7:49-62.e8.PubMedPubMedCentralCrossRef Tanner LB, Goglia AG, Wei MH, Sehgal T, Parsons LR, Park JO, et al. Four key steps control glycolytic flux in mammalian cells. Cell Syst. 2018;7:49-62.e8.PubMedPubMedCentralCrossRef
101.
102.
Zurück zum Zitat Xu D, Shao F, Bian X, Meng Y, Liang T, Lu Z. The evolving landscape of noncanonical functions of metabolic enzymes in cancer and other pathologies. Cell Metab. 2021;33:33–50.PubMedCrossRef Xu D, Shao F, Bian X, Meng Y, Liang T, Lu Z. The evolving landscape of noncanonical functions of metabolic enzymes in cancer and other pathologies. Cell Metab. 2021;33:33–50.PubMedCrossRef
103.
Zurück zum Zitat Gao X, Wang H, Yang JJ, Liu X, Liu Z-R. Pyruvate kinase M2 regulates gene transcription by acting as a protein kinase. Mol Cell. 2012;45:598–609.PubMedPubMedCentralCrossRef Gao X, Wang H, Yang JJ, Liu X, Liu Z-R. Pyruvate kinase M2 regulates gene transcription by acting as a protein kinase. Mol Cell. 2012;45:598–609.PubMedPubMedCentralCrossRef
104.
Zurück zum Zitat Li X, Jiang Y, Meisenhelder J, Yang W, Hawke DH, Zheng Y, et al. Mitochondria-translocated PGK1 functions as a protein kinase to coordinate glycolysis and the TCA cycle in tumorigenesis. Mol Cell. 2016;61:705–19.PubMedPubMedCentralCrossRef Li X, Jiang Y, Meisenhelder J, Yang W, Hawke DH, Zheng Y, et al. Mitochondria-translocated PGK1 functions as a protein kinase to coordinate glycolysis and the TCA cycle in tumorigenesis. Mol Cell. 2016;61:705–19.PubMedPubMedCentralCrossRef
105.
Zurück zum Zitat Son J, Lyssiotis CA, Ying H, Wang X, Hua S, Ligorio M, et al. Glutamine supports pancreatic cancer growth through a KRAS-regulated metabolic pathway. Nature. 2013;496:101–5.PubMedPubMedCentralCrossRef Son J, Lyssiotis CA, Ying H, Wang X, Hua S, Ligorio M, et al. Glutamine supports pancreatic cancer growth through a KRAS-regulated metabolic pathway. Nature. 2013;496:101–5.PubMedPubMedCentralCrossRef
106.
Zurück zum Zitat Ying H, Kimmelman AC, Lyssiotis CA, Hua S, Chu GC, Fletcher-Sananikone E, et al. Oncogenic Kras maintains pancreatic tumors through regulation of anabolic glucose metabolism. Cell. 2012;149:656–70.PubMedPubMedCentralCrossRef Ying H, Kimmelman AC, Lyssiotis CA, Hua S, Chu GC, Fletcher-Sananikone E, et al. Oncogenic Kras maintains pancreatic tumors through regulation of anabolic glucose metabolism. Cell. 2012;149:656–70.PubMedPubMedCentralCrossRef
108.
Zurück zum Zitat Hatina J, Kripnerova M, Houfkova K, Pesta M, Kuncova J, Sana J, et al. Sarcoma stem cell heterogeneity. In: Birbrair A, editor., et al., Stem cells heterogeneity—novel concepts. Cham: Springer International Publishing; 2019. p. 95–118.CrossRef Hatina J, Kripnerova M, Houfkova K, Pesta M, Kuncova J, Sana J, et al. Sarcoma stem cell heterogeneity. In: Birbrair A, editor., et al., Stem cells heterogeneity—novel concepts. Cham: Springer International Publishing; 2019. p. 95–118.CrossRef
109.
110.
Zurück zum Zitat Martins-Neves SR, Corver WE, Paiva-Oliveira DI, van den Akker BEWM, Briaire-de-Bruijn IH, Bovée JVMG, et al. Osteosarcoma stem cells have active Wnt/β-catenin and overexpress SOX2 and KLF4. J Cell Physiol. 2016;231:876–86.PubMedCrossRef Martins-Neves SR, Corver WE, Paiva-Oliveira DI, van den Akker BEWM, Briaire-de-Bruijn IH, Bovée JVMG, et al. Osteosarcoma stem cells have active Wnt/β-catenin and overexpress SOX2 and KLF4. J Cell Physiol. 2016;231:876–86.PubMedCrossRef
112.
Zurück zum Zitat Harvey KF, Zhang X, Thomas DM. The Hippo pathway and human cancer. Nat Rev Cancer. 2013;13:246–57.PubMedCrossRef Harvey KF, Zhang X, Thomas DM. The Hippo pathway and human cancer. Nat Rev Cancer. 2013;13:246–57.PubMedCrossRef
113.
Zurück zum Zitat Kapoor A, Yao W, Ying H, Hua S, Liewen A, Wang Q, et al. Yap1 Activation enables bypass of oncogenic Kras addiction in pancreatic cancer. Cell. 2014;158:185–97.PubMedPubMedCentralCrossRef Kapoor A, Yao W, Ying H, Hua S, Liewen A, Wang Q, et al. Yap1 Activation enables bypass of oncogenic Kras addiction in pancreatic cancer. Cell. 2014;158:185–97.PubMedPubMedCentralCrossRef
114.
Zurück zum Zitat St John MA, Tao W, Fei X, Fukumoto R, Carcangiu ML, Brownstein DG, et al. Mice deficient of Lats1 develop soft-tissue sarcomas, ovarian tumours and pituitary dysfunction. Nat Genet. 1999;21:182–6.PubMedCrossRef St John MA, Tao W, Fei X, Fukumoto R, Carcangiu ML, Brownstein DG, et al. Mice deficient of Lats1 develop soft-tissue sarcomas, ovarian tumours and pituitary dysfunction. Nat Genet. 1999;21:182–6.PubMedCrossRef
115.
Zurück zum Zitat Zhou D, Conrad C, Xia F, Park J-S, Payer B, Yin Y, et al. Mst1 and Mst2 maintain hepatocyte quiescence and suppress the development of hepatocellular carcinoma through inactivation of the Yap1 oncogene. Cancer Cell. 2009;16:425–38.PubMedPubMedCentralCrossRef Zhou D, Conrad C, Xia F, Park J-S, Payer B, Yin Y, et al. Mst1 and Mst2 maintain hepatocyte quiescence and suppress the development of hepatocellular carcinoma through inactivation of the Yap1 oncogene. Cancer Cell. 2009;16:425–38.PubMedPubMedCentralCrossRef
116.
Zurück zum Zitat Mohamed AD, Shah N, Hettmer S, Vargesson N, Wackerhage H. Analysis of the relationship between the KRAS G12V oncogene and the Hippo effector YAP1 in embryonal rhabdomyosarcoma. Sci Rep. 2018;8:15674.PubMedPubMedCentralCrossRef Mohamed AD, Shah N, Hettmer S, Vargesson N, Wackerhage H. Analysis of the relationship between the KRAS G12V oncogene and the Hippo effector YAP1 in embryonal rhabdomyosarcoma. Sci Rep. 2018;8:15674.PubMedPubMedCentralCrossRef
117.
118.
Zurück zum Zitat Santinon G, Pocaterra A, Dupont S. Control of YAP/TAZ activity by metabolic and nutrient-sensing pathways. Trends Cell Biol. 2016;26:289–99.PubMedCrossRef Santinon G, Pocaterra A, Dupont S. Control of YAP/TAZ activity by metabolic and nutrient-sensing pathways. Trends Cell Biol. 2016;26:289–99.PubMedCrossRef
119.
Zurück zum Zitat Wang W, Xiao Z-D, Li X, Aziz KE, Gan B, Johnson RL, et al. AMPK modulates Hippo pathway activity to regulate energy homeostasis. Nat Cell Biol. 2015;17:490–9.PubMedPubMedCentralCrossRef Wang W, Xiao Z-D, Li X, Aziz KE, Gan B, Johnson RL, et al. AMPK modulates Hippo pathway activity to regulate energy homeostasis. Nat Cell Biol. 2015;17:490–9.PubMedPubMedCentralCrossRef
120.
Zurück zum Zitat Enzo E, Santinon G, Pocaterra A, Aragona M, Bresolin S, Forcato M, et al. Aerobic glycolysis tunes YAP/TAZ transcriptional activity. EMBO J. 2015;34:1349–70.PubMedPubMedCentralCrossRef Enzo E, Santinon G, Pocaterra A, Aragona M, Bresolin S, Forcato M, et al. Aerobic glycolysis tunes YAP/TAZ transcriptional activity. EMBO J. 2015;34:1349–70.PubMedPubMedCentralCrossRef
121.
Zurück zum Zitat Mota MSV, Jackson WP, Bailey SK, Vayalil P, Landar A, Rostas JW, et al. Deficiency of tumor suppressor Merlin facilitates metabolic adaptation by co-operative engagement of SMAD-Hippo signaling in breast cancer. Carcinogenesis. 2018;39:1165–75.PubMedPubMedCentralCrossRef Mota MSV, Jackson WP, Bailey SK, Vayalil P, Landar A, Rostas JW, et al. Deficiency of tumor suppressor Merlin facilitates metabolic adaptation by co-operative engagement of SMAD-Hippo signaling in breast cancer. Carcinogenesis. 2018;39:1165–75.PubMedPubMedCentralCrossRef
122.
Zurück zum Zitat Agresta L, Salloum R, Hummel TR, Ratner N, Mangano FT, Fuller C, et al. Malignant peripheral nerve sheath tumor: transformation in a patient with neurofibromatosis type 2. Pediatr Blood Cancer. 2019;66:e27520.PubMedCrossRef Agresta L, Salloum R, Hummel TR, Ratner N, Mangano FT, Fuller C, et al. Malignant peripheral nerve sheath tumor: transformation in a patient with neurofibromatosis type 2. Pediatr Blood Cancer. 2019;66:e27520.PubMedCrossRef
123.
Zurück zum Zitat White SM, Avantaggiati ML, Nemazanyy I, Di Poto C, Yang Y, Pende M, et al. YAP/TAZ inhibition induces metabolic and signaling rewiring resulting in targetable vulnerabilities in NF2-deficient tumor cells. Dev Cell. 2019;49:425-443.e9.PubMedPubMedCentralCrossRef White SM, Avantaggiati ML, Nemazanyy I, Di Poto C, Yang Y, Pende M, et al. YAP/TAZ inhibition induces metabolic and signaling rewiring resulting in targetable vulnerabilities in NF2-deficient tumor cells. Dev Cell. 2019;49:425-443.e9.PubMedPubMedCentralCrossRef
124.
Zurück zum Zitat Chen R, Zhu S, Fan X-G, Wang H, Lotze MT, Zeh HJ, et al. High mobility group protein B1 controls liver cancer initiation through yes-associated protein-dependent aerobic glycolysis. Hepatology. 2018;67:1823–41.PubMedCrossRef Chen R, Zhu S, Fan X-G, Wang H, Lotze MT, Zeh HJ, et al. High mobility group protein B1 controls liver cancer initiation through yes-associated protein-dependent aerobic glycolysis. Hepatology. 2018;67:1823–41.PubMedCrossRef
125.
Zurück zum Zitat Zhang X, Li Y, Ma Y, Yang L, Wang T, Meng X, et al. Yes-associated protein (YAP) binds to HIF-1α and sustains HIF-1α protein stability to promote hepatocellular carcinoma cell glycolysis under hypoxic stress. J Exp Clin Cancer Res. 2018;37:216.PubMedPubMedCentralCrossRef Zhang X, Li Y, Ma Y, Yang L, Wang T, Meng X, et al. Yes-associated protein (YAP) binds to HIF-1α and sustains HIF-1α protein stability to promote hepatocellular carcinoma cell glycolysis under hypoxic stress. J Exp Clin Cancer Res. 2018;37:216.PubMedPubMedCentralCrossRef
126.
Zurück zum Zitat Rivera-Reyes A, Ye S, Marino GE, Egolf S, Ciotti GE, Chor S, et al. YAP1 enhances NF-κB-dependent and independent effects on clock-mediated unfolded protein responses and autophagy in sarcoma. Cell Death Dis. 2018;9:1108.PubMedPubMedCentralCrossRef Rivera-Reyes A, Ye S, Marino GE, Egolf S, Ciotti GE, Chor S, et al. YAP1 enhances NF-κB-dependent and independent effects on clock-mediated unfolded protein responses and autophagy in sarcoma. Cell Death Dis. 2018;9:1108.PubMedPubMedCentralCrossRef
128.
Zurück zum Zitat Birsoy K, Wang T, Chen WW, Freinkman E, Abu-Remaileh M, Sabatini DM. An essential role of the mitochondrial electron transport chain in cell proliferation is to enable aspartate synthesis. Cell. 2015;162:540–51.PubMedPubMedCentralCrossRef Birsoy K, Wang T, Chen WW, Freinkman E, Abu-Remaileh M, Sabatini DM. An essential role of the mitochondrial electron transport chain in cell proliferation is to enable aspartate synthesis. Cell. 2015;162:540–51.PubMedPubMedCentralCrossRef
129.
Zurück zum Zitat Sullivan LB, Gui DY, Hosios AM, Bush LN, Freinkman E, Vander Heiden MG. Supporting aspartate biosynthesis is an essential function of respiration in proliferating cells. Cell. 2015;162:552–63.PubMedPubMedCentralCrossRef Sullivan LB, Gui DY, Hosios AM, Bush LN, Freinkman E, Vander Heiden MG. Supporting aspartate biosynthesis is an essential function of respiration in proliferating cells. Cell. 2015;162:552–63.PubMedPubMedCentralCrossRef
130.
Zurück zum Zitat Ma Q, Cavallin LE, Leung HJ, Chiozzini C, Goldschmidt-Clermont PJ, Mesri EA. A role for virally induced reactive oxygen species in Kaposi’s sarcoma herpesvirus tumorigenesis. Antioxid Redox Signal. 2013;18:80–90.PubMedPubMedCentralCrossRef Ma Q, Cavallin LE, Leung HJ, Chiozzini C, Goldschmidt-Clermont PJ, Mesri EA. A role for virally induced reactive oxygen species in Kaposi’s sarcoma herpesvirus tumorigenesis. Antioxid Redox Signal. 2013;18:80–90.PubMedPubMedCentralCrossRef
131.
Zurück zum Zitat Ma T, Patel H, Babapoor-Farrokhran S, Franklin R, Semenza GL, Sodhi A, et al. KSHV induces aerobic glycolysis and angiogenesis through HIF-1-dependent upregulation of pyruvate kinase 2 in Kaposi’s sarcoma. Angiogenesis. 2015;18:477–88.PubMedPubMedCentralCrossRef Ma T, Patel H, Babapoor-Farrokhran S, Franklin R, Semenza GL, Sodhi A, et al. KSHV induces aerobic glycolysis and angiogenesis through HIF-1-dependent upregulation of pyruvate kinase 2 in Kaposi’s sarcoma. Angiogenesis. 2015;18:477–88.PubMedPubMedCentralCrossRef
132.
Zurück zum Zitat Delgado T, Carroll PA, Punjabi AS, Margineantu D, Hockenbery DM, Lagunoff M. Induction of the Warburg effect by Kaposi’s sarcoma herpesvirus is required for the maintenance of latently infected endothelial cells. Proc Natl Acad Sci. 2010;107:10696–701.PubMedPubMedCentralCrossRef Delgado T, Carroll PA, Punjabi AS, Margineantu D, Hockenbery DM, Lagunoff M. Induction of the Warburg effect by Kaposi’s sarcoma herpesvirus is required for the maintenance of latently infected endothelial cells. Proc Natl Acad Sci. 2010;107:10696–701.PubMedPubMedCentralCrossRef
133.
Zurück zum Zitat Cavallin LE, Ma Q, Naipauer J, Gupta S, Kurian M, Locatelli P, et al. KSHV-induced ligand mediated activation of PDGF receptor-alpha drives Kaposi’s sarcomagenesis. PLoS Pathog. 2018;14:e1007175.PubMedPubMedCentralCrossRef Cavallin LE, Ma Q, Naipauer J, Gupta S, Kurian M, Locatelli P, et al. KSHV-induced ligand mediated activation of PDGF receptor-alpha drives Kaposi’s sarcomagenesis. PLoS Pathog. 2018;14:e1007175.PubMedPubMedCentralCrossRef
134.
Zurück zum Zitat Veeranna RP, Haque M, Davis DA, Yang M, Yarchoan R. Kaposi’s sarcoma-associated herpesvirus latency-associated nuclear antigen induction by hypoxia and hypoxia-inducible factors. J Virol. 2012;86:1097–108.PubMedPubMedCentralCrossRef Veeranna RP, Haque M, Davis DA, Yang M, Yarchoan R. Kaposi’s sarcoma-associated herpesvirus latency-associated nuclear antigen induction by hypoxia and hypoxia-inducible factors. J Virol. 2012;86:1097–108.PubMedPubMedCentralCrossRef
135.
Zurück zum Zitat Choi UY, Lee JJ, Park A, Zhu W, Lee H-R, Choi YJ, et al. Oncogenic human herpesvirus hijacks proline metabolism for tumorigenesis. PNAS. 2020;117:8083–93.PubMedPubMedCentralCrossRef Choi UY, Lee JJ, Park A, Zhu W, Lee H-R, Choi YJ, et al. Oncogenic human herpesvirus hijacks proline metabolism for tumorigenesis. PNAS. 2020;117:8083–93.PubMedPubMedCentralCrossRef
136.
Zurück zum Zitat Locasale JW, Grassian AR, Melman T, Lyssiotis CA, Mattaini KR, Bass AJ, et al. Phosphoglycerate dehydrogenase diverts glycolytic flux and contributes to oncogenesis. Nat Genet. 2011;43:869–74.PubMedPubMedCentralCrossRef Locasale JW, Grassian AR, Melman T, Lyssiotis CA, Mattaini KR, Bass AJ, et al. Phosphoglycerate dehydrogenase diverts glycolytic flux and contributes to oncogenesis. Nat Genet. 2011;43:869–74.PubMedPubMedCentralCrossRef
138.
Zurück zum Zitat Gao X, Sanderson SM, Dai Z, Reid MA, Cooper DE, Lu M, et al. Dietary methionine influences therapy in mouse cancer models and alters human metabolism. Nature. 2019;572:397–401.PubMedPubMedCentralCrossRef Gao X, Sanderson SM, Dai Z, Reid MA, Cooper DE, Lu M, et al. Dietary methionine influences therapy in mouse cancer models and alters human metabolism. Nature. 2019;572:397–401.PubMedPubMedCentralCrossRef
139.
Zurück zum Zitat Nusse R, Clevers H. Wnt/β-catenin signaling, disease, and emerging therapeutic modalities. Cell. 2017;169:985–99.PubMedCrossRef Nusse R, Clevers H. Wnt/β-catenin signaling, disease, and emerging therapeutic modalities. Cell. 2017;169:985–99.PubMedCrossRef
140.
Zurück zum Zitat Esen E, Chen J, Karner CM, Okunade AL, Patterson BW, Long F. WNT-LRP5 signaling induces Warburg effect through mTORC2 activation during osteoblast differentiation. Cell Metab. 2013;17:745–55.PubMedPubMedCentralCrossRef Esen E, Chen J, Karner CM, Okunade AL, Patterson BW, Long F. WNT-LRP5 signaling induces Warburg effect through mTORC2 activation during osteoblast differentiation. Cell Metab. 2013;17:745–55.PubMedPubMedCentralCrossRef
141.
Zurück zum Zitat Karner CM, Esen E, Chen J, Hsu F-F, Turk J, Long F. Wnt protein signaling reduces nuclear acetyl-CoA levels to suppress gene expression during osteoblast differentiation. J Biol Chem. 2016;291:13028–39.PubMedPubMedCentralCrossRef Karner CM, Esen E, Chen J, Hsu F-F, Turk J, Long F. Wnt protein signaling reduces nuclear acetyl-CoA levels to suppress gene expression during osteoblast differentiation. J Biol Chem. 2016;291:13028–39.PubMedPubMedCentralCrossRef
142.
Zurück zum Zitat Frey JL, Li Z, Ellis JM, Zhang Q, Farber CR, Aja S, et al. Wnt-Lrp5 signaling regulates fatty acid metabolism in the osteoblast. Mol Cell Biol. 2015;35:1979–91.PubMedPubMedCentralCrossRef Frey JL, Li Z, Ellis JM, Zhang Q, Farber CR, Aja S, et al. Wnt-Lrp5 signaling regulates fatty acid metabolism in the osteoblast. Mol Cell Biol. 2015;35:1979–91.PubMedPubMedCentralCrossRef
143.
Zurück zum Zitat Danieau G, Morice S, Rédini F, Verrecchia F, Royer BL. New insights about the Wnt/β-catenin signaling pathway in primary bone tumors and their microenvironment: a promising target to develop therapeutic strategies? IJMS. 2019;20:3751.PubMedCentralCrossRef Danieau G, Morice S, Rédini F, Verrecchia F, Royer BL. New insights about the Wnt/β-catenin signaling pathway in primary bone tumors and their microenvironment: a promising target to develop therapeutic strategies? IJMS. 2019;20:3751.PubMedCentralCrossRef
144.
Zurück zum Zitat Fuxe J, Vincent T, Garcia de Herreros A. Transcriptional crosstalk between TGFβ and stem cell pathways in tumor cell invasion: role of EMT promoting Smad complexes. Cell Cycle. 2010;9:2363–74.PubMedCrossRef Fuxe J, Vincent T, Garcia de Herreros A. Transcriptional crosstalk between TGFβ and stem cell pathways in tumor cell invasion: role of EMT promoting Smad complexes. Cell Cycle. 2010;9:2363–74.PubMedCrossRef
145.
Zurück zum Zitat Briski LM, Thomas DG, Patel RM, Lawlor ER, Chugh R, McHugh JB, et al. Canonical Wnt/β-catenin signaling activation in soft-tissue sarcomas: a comparative study of synovial sarcoma and leiomyosarcoma. Rare Tumors. 2018;10:203636131881343.CrossRef Briski LM, Thomas DG, Patel RM, Lawlor ER, Chugh R, McHugh JB, et al. Canonical Wnt/β-catenin signaling activation in soft-tissue sarcomas: a comparative study of synovial sarcoma and leiomyosarcoma. Rare Tumors. 2018;10:203636131881343.CrossRef
146.
147.
Zurück zum Zitat Wang W-G, Chen S-J, He J-S, Li J-S, Zang X-F. The tumor suppressive role of RASSF1A in osteosarcoma through the Wnt signaling pathway. Tumor Biol. 2016;37:8869–77.CrossRef Wang W-G, Chen S-J, He J-S, Li J-S, Zang X-F. The tumor suppressive role of RASSF1A in osteosarcoma through the Wnt signaling pathway. Tumor Biol. 2016;37:8869–77.CrossRef
148.
Zurück zum Zitat Li Y, Zhang S, Zhang C, Wang M. LncRNA MEG3 inhibits the inflammatory response of ankylosing spondylitis by targeting miR-146a. Mol Cell Biochem. 2020;466:17–24.PubMedCrossRef Li Y, Zhang S, Zhang C, Wang M. LncRNA MEG3 inhibits the inflammatory response of ankylosing spondylitis by targeting miR-146a. Mol Cell Biochem. 2020;466:17–24.PubMedCrossRef
149.
Zurück zum Zitat Techavichit P, Gao Y, Kurenbekova L, Shuck R, Donehower LA, Yustein JT. Secreted Frizzled-Related Protein 2 (sFRP2) promotes osteosarcoma invasion and metastatic potential. BMC Cancer. 2016;16:869.PubMedPubMedCentralCrossRef Techavichit P, Gao Y, Kurenbekova L, Shuck R, Donehower LA, Yustein JT. Secreted Frizzled-Related Protein 2 (sFRP2) promotes osteosarcoma invasion and metastatic potential. BMC Cancer. 2016;16:869.PubMedPubMedCentralCrossRef
150.
Zurück zum Zitat Jacks T, Remington L, Williams BO, Schmitt EM, Halachmi S, Bronson RT, et al. Tumor spectrum analysis in p53-mutant mice. Curr Biol. 1994;4:1–7.PubMedCrossRef Jacks T, Remington L, Williams BO, Schmitt EM, Halachmi S, Bronson RT, et al. Tumor spectrum analysis in p53-mutant mice. Curr Biol. 1994;4:1–7.PubMedCrossRef
151.
Zurück zum Zitat Barretina J, Taylor BS, Banerji S, Ramos AH, Lagos-Quintana M, DeCarolis PL, et al. Subtype-specific genomic alterations define new targets for soft-tissue sarcoma therapy. Nat Genet. 2010;42:715–21.PubMedPubMedCentralCrossRef Barretina J, Taylor BS, Banerji S, Ramos AH, Lagos-Quintana M, DeCarolis PL, et al. Subtype-specific genomic alterations define new targets for soft-tissue sarcoma therapy. Nat Genet. 2010;42:715–21.PubMedPubMedCentralCrossRef
152.
Zurück zum Zitat Blay J-Y, Ray-Coquard I. Evolving biological understanding and treatment of sarcomas. Nat Rev Clin Oncol. 2017;14:78–80.PubMedCrossRef Blay J-Y, Ray-Coquard I. Evolving biological understanding and treatment of sarcomas. Nat Rev Clin Oncol. 2017;14:78–80.PubMedCrossRef
153.
Zurück zum Zitat Bui NQ. A clinico-genomic analysis of soft tissue sarcoma patients reveals CDKN2A deletion as a biomarker for poor prognosis. Clin Sarcoma Res. 2019;9:11.CrossRef Bui NQ. A clinico-genomic analysis of soft tissue sarcoma patients reveals CDKN2A deletion as a biomarker for poor prognosis. Clin Sarcoma Res. 2019;9:11.CrossRef
154.
Zurück zum Zitat Mito JK, Riedel RF, Dodd L, Lahat G, Lazar AJ, Dodd RD, et al. Cross species genomic analysis identifies a mouse model as undifferentiated pleomorphic sarcoma/malignant fibrous histiocytoma. PLoS ONE. 2009;4:e8075.PubMedPubMedCentralCrossRef Mito JK, Riedel RF, Dodd L, Lahat G, Lazar AJ, Dodd RD, et al. Cross species genomic analysis identifies a mouse model as undifferentiated pleomorphic sarcoma/malignant fibrous histiocytoma. PLoS ONE. 2009;4:e8075.PubMedPubMedCentralCrossRef
155.
Zurück zum Zitat Kirsch DG, Dinulescu DM, Miller JB, Grimm J, Santiago PM, Young NP, et al. A spatially and temporally restricted mouse model of soft tissue sarcoma. Nat Med. 2007;13:992–7.PubMedCrossRef Kirsch DG, Dinulescu DM, Miller JB, Grimm J, Santiago PM, Young NP, et al. A spatially and temporally restricted mouse model of soft tissue sarcoma. Nat Med. 2007;13:992–7.PubMedCrossRef
156.
Zurück zum Zitat Liu J, Zhang C, Hu W, Feng Z. Tumor suppressor p53 and its mutants in cancer metabolism. Cancer Lett. 2015;356:197–203.PubMedCrossRef Liu J, Zhang C, Hu W, Feng Z. Tumor suppressor p53 and its mutants in cancer metabolism. Cancer Lett. 2015;356:197–203.PubMedCrossRef
157.
Zurück zum Zitat Simabuco FM, Morale MG, Pavan ICB, Morelli AP, Silva FR, Tamura RE. p53 and metabolism: from mechanism to therapeutics. Oncotarget. 2018;9:23780–823.PubMedPubMedCentralCrossRef Simabuco FM, Morale MG, Pavan ICB, Morelli AP, Silva FR, Tamura RE. p53 and metabolism: from mechanism to therapeutics. Oncotarget. 2018;9:23780–823.PubMedPubMedCentralCrossRef
158.
Zurück zum Zitat Valente LJ, Gray DHD, Michalak EM, Pinon-Hofbauer J, Egle A, Scott CL, et al. p53 efficiently suppresses tumor development in the complete absence of its cell-cycle inhibitory and proapoptotic effectors p21, Puma, and Noxa. Cell Rep. 2013;3:1339–45.PubMedCrossRef Valente LJ, Gray DHD, Michalak EM, Pinon-Hofbauer J, Egle A, Scott CL, et al. p53 efficiently suppresses tumor development in the complete absence of its cell-cycle inhibitory and proapoptotic effectors p21, Puma, and Noxa. Cell Rep. 2013;3:1339–45.PubMedCrossRef
160.
Zurück zum Zitat Ahrens WA, Ridenour RV, Caron BL, Miller DV, Folpe AL. GLUT-1 expression in mesenchymal tumors: an immunohistochemical study of 247 soft tissue and bone neoplasms. Hum Pathol. 2008;39:1519–26.PubMedCrossRef Ahrens WA, Ridenour RV, Caron BL, Miller DV, Folpe AL. GLUT-1 expression in mesenchymal tumors: an immunohistochemical study of 247 soft tissue and bone neoplasms. Hum Pathol. 2008;39:1519–26.PubMedCrossRef
161.
Zurück zum Zitat Schwartzenberg-Bar-Yoseph F, Armoni M, Karnieli E. The tumor suppressor p53 down-regulates glucose transporters GLUT1 and GLUT4 gene expression. Cancer Res. 2004;64:2627–33.PubMedCrossRef Schwartzenberg-Bar-Yoseph F, Armoni M, Karnieli E. The tumor suppressor p53 down-regulates glucose transporters GLUT1 and GLUT4 gene expression. Cancer Res. 2004;64:2627–33.PubMedCrossRef
162.
Zurück zum Zitat Bensaad K, Tsuruta A, Selak MA, Vidal MNC, Nakano K, Bartrons R, et al. TIGAR, a p53-inducible regulator of glycolysis and apoptosis. Cell. 2006;126:107–20.PubMedCrossRef Bensaad K, Tsuruta A, Selak MA, Vidal MNC, Nakano K, Bartrons R, et al. TIGAR, a p53-inducible regulator of glycolysis and apoptosis. Cell. 2006;126:107–20.PubMedCrossRef
163.
Zurück zum Zitat Zhang C, Lin M, Wu R, Wang X, Yang B, Levine AJ, et al. Parkin, a p53 target gene, mediates the role of p53 in glucose metabolism and the Warburg effect. Proc Natl Acad Sci. 2011;108:16259–64.PubMedPubMedCentralCrossRef Zhang C, Lin M, Wu R, Wang X, Yang B, Levine AJ, et al. Parkin, a p53 target gene, mediates the role of p53 in glucose metabolism and the Warburg effect. Proc Natl Acad Sci. 2011;108:16259–64.PubMedPubMedCentralCrossRef
164.
Zurück zum Zitat Jiang P, Du W, Wang X, Mancuso A, Gao X, Wu M, et al. p53 regulates biosynthesis through direct inactivation of glucose-6-phosphate dehydrogenase. Nat Cell Biol. 2011;13:15.CrossRef Jiang P, Du W, Wang X, Mancuso A, Gao X, Wu M, et al. p53 regulates biosynthesis through direct inactivation of glucose-6-phosphate dehydrogenase. Nat Cell Biol. 2011;13:15.CrossRef
165.
Zurück zum Zitat Erpapazoglou Z, Corti O. The endoplasmic reticulum/mitochondria interface: a subcellular platform for the orchestration of the functions of the PINK1–Parkin pathway? Biochem Soc Trans. 2015;43:297–301.PubMedCrossRef Erpapazoglou Z, Corti O. The endoplasmic reticulum/mitochondria interface: a subcellular platform for the orchestration of the functions of the PINK1–Parkin pathway? Biochem Soc Trans. 2015;43:297–301.PubMedCrossRef
166.
Zurück zum Zitat Kruiswijk F, Labuschagne CF, Vousden KH. p53 in survival, death and metabolic health: a lifeguard with a licence to kill. Nat Rev Mol Cell Biol. 2015;16:393–405.PubMedCrossRef Kruiswijk F, Labuschagne CF, Vousden KH. p53 in survival, death and metabolic health: a lifeguard with a licence to kill. Nat Rev Mol Cell Biol. 2015;16:393–405.PubMedCrossRef
167.
Zurück zum Zitat Morris JP, Yashinskie JJ, Koche R, Chandwani R, Tian S, Chen C-C, et al. α-Ketoglutarate links p53 to cell fate during tumour suppression. Nature. 2019;573:595–9.PubMedPubMedCentralCrossRef Morris JP, Yashinskie JJ, Koche R, Chandwani R, Tian S, Chen C-C, et al. α-Ketoglutarate links p53 to cell fate during tumour suppression. Nature. 2019;573:595–9.PubMedPubMedCentralCrossRef
168.
Zurück zum Zitat Gaude E, Frezza C. Tissue-specific and convergent metabolic transformation of cancer correlates with metastatic potential and patient survival. Nat Commun. 2016;7:13041.PubMedPubMedCentralCrossRef Gaude E, Frezza C. Tissue-specific and convergent metabolic transformation of cancer correlates with metastatic potential and patient survival. Nat Commun. 2016;7:13041.PubMedPubMedCentralCrossRef
169.
Zurück zum Zitat Benz MR, Tchekmedyian N, Eilber FC, Federman N, Czernin J, Tap WD. Utilization of positron emission tomography in the management of patients with sarcoma. Curr Opin Oncol. 2009;21:345–51.PubMedCrossRef Benz MR, Tchekmedyian N, Eilber FC, Federman N, Czernin J, Tap WD. Utilization of positron emission tomography in the management of patients with sarcoma. Curr Opin Oncol. 2009;21:345–51.PubMedCrossRef
170.
Zurück zum Zitat Wagner LM, Kremer N, Gelfand MJ, Sharp SE, Turpin BK, Nagarajan R, et al. Detection of lymph node metastases in pediatric and adolescent/young adult sarcoma: sentinel lymph node biopsy versus fludeoxyglucose positron emission tomography imaging—a prospective trial: sentinel lymph node biopsy versus PET in sarcoma. Cancer. 2017;123:155–60.PubMedCrossRef Wagner LM, Kremer N, Gelfand MJ, Sharp SE, Turpin BK, Nagarajan R, et al. Detection of lymph node metastases in pediatric and adolescent/young adult sarcoma: sentinel lymph node biopsy versus fludeoxyglucose positron emission tomography imaging—a prospective trial: sentinel lymph node biopsy versus PET in sarcoma. Cancer. 2017;123:155–60.PubMedCrossRef
171.
Zurück zum Zitat Issaq SH, Teicher BA, Monks A. Bioenergetic properties of human sarcoma cells help define sensitivity to metabolic inhibitors. Cell Cycle. 2014;13:1152–61.PubMedPubMedCentralCrossRef Issaq SH, Teicher BA, Monks A. Bioenergetic properties of human sarcoma cells help define sensitivity to metabolic inhibitors. Cell Cycle. 2014;13:1152–61.PubMedPubMedCentralCrossRef
172.
Zurück zum Zitat Dasgupta A, Trucco M, Rainusso N, Bernardi RJ, Shuck R, Kurenbekova L, et al. Metabolic modulation of Ewing sarcoma cells inhibits tumor growth and stem cell properties. Oncotarget. 2017;8:77292–308.PubMedPubMedCentralCrossRef Dasgupta A, Trucco M, Rainusso N, Bernardi RJ, Shuck R, Kurenbekova L, et al. Metabolic modulation of Ewing sarcoma cells inhibits tumor growth and stem cell properties. Oncotarget. 2017;8:77292–308.PubMedPubMedCentralCrossRef
173.
Zurück zum Zitat Mao L, Dauchy RT, Blask DE, Dauchy EM, Slakey LM, Brimer S, et al. Melatonin suppression of aerobic glycolysis (Warburg effect), survival signalling and metastasis in human leiomyosarcoma. J Pineal Res. 2016;60:167–77.PubMedCrossRef Mao L, Dauchy RT, Blask DE, Dauchy EM, Slakey LM, Brimer S, et al. Melatonin suppression of aerobic glycolysis (Warburg effect), survival signalling and metastasis in human leiomyosarcoma. J Pineal Res. 2016;60:167–77.PubMedCrossRef
174.
Zurück zum Zitat Li B, Qiu B, Lee DSM, Walton ZE, Ochocki JD, Mathew LK, et al. Fructose-1,6-bisphosphatase opposes renal carcinoma progression. Nature. 2014;513:251–5.PubMedPubMedCentralCrossRef Li B, Qiu B, Lee DSM, Walton ZE, Ochocki JD, Mathew LK, et al. Fructose-1,6-bisphosphatase opposes renal carcinoma progression. Nature. 2014;513:251–5.PubMedPubMedCentralCrossRef
175.
Zurück zum Zitat Issaq SH, Mendoza A, Fox SD, Helman LJ. Glutamine synthetase is necessary for sarcoma adaptation to glutamine deprivation and tumor growth. Oncogenesis. 2019;8:20.PubMedPubMedCentralCrossRef Issaq SH, Mendoza A, Fox SD, Helman LJ. Glutamine synthetase is necessary for sarcoma adaptation to glutamine deprivation and tumor growth. Oncogenesis. 2019;8:20.PubMedPubMedCentralCrossRef
176.
Zurück zum Zitat Han J, Zhang Y, Xu J, Zhang T, Wang H, Wang Z, et al. Her4 promotes cancer metabolic reprogramming via the c-Myc-dependent signaling axis. Cancer Lett. 2021;496:57–71.PubMedCrossRef Han J, Zhang Y, Xu J, Zhang T, Wang H, Wang Z, et al. Her4 promotes cancer metabolic reprogramming via the c-Myc-dependent signaling axis. Cancer Lett. 2021;496:57–71.PubMedCrossRef
177.
Zurück zum Zitat Masoud R, Reyes-Castellanos G, Lac S, Garcia J, Dou S, Shintu L, et al. Targeting mitochondrial complex i overcomes chemoresistance in high OXPHOS pancreatic cancer. Cell Rep Med. 2020;1:100143.PubMedPubMedCentralCrossRef Masoud R, Reyes-Castellanos G, Lac S, Garcia J, Dou S, Shintu L, et al. Targeting mitochondrial complex i overcomes chemoresistance in high OXPHOS pancreatic cancer. Cell Rep Med. 2020;1:100143.PubMedPubMedCentralCrossRef
178.
Zurück zum Zitat Mor I, Cheung EC, Vousden KH. Control of glycolysis through regulation of PFK1: old friends and recent additions. Cold Spring Harb Symp Quant Biol. 2011;76:211–6.PubMedCrossRef Mor I, Cheung EC, Vousden KH. Control of glycolysis through regulation of PFK1: old friends and recent additions. Cold Spring Harb Symp Quant Biol. 2011;76:211–6.PubMedCrossRef
179.
Zurück zum Zitat Moreno-Sánchez R, Marín-Hernández A, Gallardo-Pérez JC, Quezada H, Encalada R, Rodríguez-Enríquez S, et al. Phosphofructokinase type 1 kinetics, isoform expression, and gene polymorphisms in cancer cells. J Cell Biochem. 2012;113:1692–703.PubMed Moreno-Sánchez R, Marín-Hernández A, Gallardo-Pérez JC, Quezada H, Encalada R, Rodríguez-Enríquez S, et al. Phosphofructokinase type 1 kinetics, isoform expression, and gene polymorphisms in cancer cells. J Cell Biochem. 2012;113:1692–703.PubMed
180.
Zurück zum Zitat Cabrera R, Baez M, Pereira HM. Kinetic and structural analysis of the allosteric ATP inhibition S. J Biol Chem. 2011;286:11. Cabrera R, Baez M, Pereira HM. Kinetic and structural analysis of the allosteric ATP inhibition S. J Biol Chem. 2011;286:11.
181.
Zurück zum Zitat Yalcin A, Telang S, Clem B, Chesney J. Regulation of glucose metabolism by 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatases in cancer. Exp Mol Pathol. 2009;86:174–9.PubMedCrossRef Yalcin A, Telang S, Clem B, Chesney J. Regulation of glucose metabolism by 6-phosphofructo-2-kinase/fructose-2,6-bisphosphatases in cancer. Exp Mol Pathol. 2009;86:174–9.PubMedCrossRef
182.
183.
Zurück zum Zitat Borst P. The malate–aspartate shuttle (Borst cycle): how it started and developed into a major metabolic pathway. IUBMB Life. 2020;72:2241–59.PubMedPubMedCentralCrossRef Borst P. The malate–aspartate shuttle (Borst cycle): how it started and developed into a major metabolic pathway. IUBMB Life. 2020;72:2241–59.PubMedPubMedCentralCrossRef
184.
Zurück zum Zitat Dai Z, Shestov AA, Lai L, Locasale JW. A flux balance of glucose metabolism clarifies the requirements of the Warburg effect. Biophys J. 2016;111:1088–100.PubMedPubMedCentralCrossRef Dai Z, Shestov AA, Lai L, Locasale JW. A flux balance of glucose metabolism clarifies the requirements of the Warburg effect. Biophys J. 2016;111:1088–100.PubMedPubMedCentralCrossRef
185.
Zurück zum Zitat Altinok O, Poggio JL, Stein DE, Bowne WB, Shieh AC, Snyder NW, et al. Malate–aspartate shuttle promotes l-lactate oxidation in mitochondria. J Cell Physiol. 2020;235:2569–81.PubMedCrossRef Altinok O, Poggio JL, Stein DE, Bowne WB, Shieh AC, Snyder NW, et al. Malate–aspartate shuttle promotes l-lactate oxidation in mitochondria. J Cell Physiol. 2020;235:2569–81.PubMedCrossRef
186.
Zurück zum Zitat Young A, Oldford C, Mailloux RJ. Lactate dehydrogenase supports lactate oxidation in mitochondria isolated from different mouse tissues. Redox Biol. 2020;28:101339.PubMedCrossRef Young A, Oldford C, Mailloux RJ. Lactate dehydrogenase supports lactate oxidation in mitochondria isolated from different mouse tissues. Redox Biol. 2020;28:101339.PubMedCrossRef
187.
Zurück zum Zitat Bonnet S, Archer SL, Allalunis-Turner J, Haromy A, Beaulieu C, Thompson R, et al. A mitochondria-K+ channel axis is suppressed in cancer and its normalization promotes apoptosis and inhibits cancer growth. Cancer Cell. 2007;11:37–51.PubMedCrossRef Bonnet S, Archer SL, Allalunis-Turner J, Haromy A, Beaulieu C, Thompson R, et al. A mitochondria-K+ channel axis is suppressed in cancer and its normalization promotes apoptosis and inhibits cancer growth. Cancer Cell. 2007;11:37–51.PubMedCrossRef
188.
189.
Zurück zum Zitat Titova E, Shagieva G, Ivanova O, Domnina L, Domninskaya M, Strelkova O, et al. Mitochondria-targeted antioxidant SkQ1 suppresses fibrosarcoma and rhabdomyosarcoma tumour cell growth. Cell Cycle. 2018;17:1797–811.PubMedPubMedCentralCrossRef Titova E, Shagieva G, Ivanova O, Domnina L, Domninskaya M, Strelkova O, et al. Mitochondria-targeted antioxidant SkQ1 suppresses fibrosarcoma and rhabdomyosarcoma tumour cell growth. Cell Cycle. 2018;17:1797–811.PubMedPubMedCentralCrossRef
190.
Zurück zum Zitat Leonardi R, Zhang Y, Rock C, Jackowski S. Coenzyme A: back in action. Prog Lipid Res. 2005;44:125–53.PubMedCrossRef Leonardi R, Zhang Y, Rock C, Jackowski S. Coenzyme A: back in action. Prog Lipid Res. 2005;44:125–53.PubMedCrossRef
191.
Zurück zum Zitat Granjeaud S, Naquet P, Galland F. An ESTs description of the new Vanin gene family conserved from fly to human. Immunogenetics. 1999;49:964–72.PubMedCrossRef Granjeaud S, Naquet P, Galland F. An ESTs description of the new Vanin gene family conserved from fly to human. Immunogenetics. 1999;49:964–72.PubMedCrossRef
192.
Zurück zum Zitat Naquet P, Kerr EW, Vickers SD, Leonardi R. Regulation of coenzyme A levels by degradation: the ‘Ins and Outs.’ Prog Lipid Res. 2020;78:101028.PubMedPubMedCentralCrossRef Naquet P, Kerr EW, Vickers SD, Leonardi R. Regulation of coenzyme A levels by degradation: the ‘Ins and Outs.’ Prog Lipid Res. 2020;78:101028.PubMedPubMedCentralCrossRef
193.
Zurück zum Zitat Scarpulla RC. Nuclear activators and coactivators in mammalian mitochondrial biogenesis. Biochim Biophys Acta (BBA) Gene Struct Expr. 2002;1576:1–14.CrossRef Scarpulla RC. Nuclear activators and coactivators in mammalian mitochondrial biogenesis. Biochim Biophys Acta (BBA) Gene Struct Expr. 2002;1576:1–14.CrossRef
194.
Zurück zum Zitat Bakkar N, Wang J, Ladner KJ, Wang H, Dahlman JM, Carathers M, et al. IKK/NF-κB regulates skeletal myogenesis via a signaling switch to inhibit differentiation and promote mitochondrial biogenesis. J Cell Biol. 2008;180:787–802.PubMedPubMedCentralCrossRef Bakkar N, Wang J, Ladner KJ, Wang H, Dahlman JM, Carathers M, et al. IKK/NF-κB regulates skeletal myogenesis via a signaling switch to inhibit differentiation and promote mitochondrial biogenesis. J Cell Biol. 2008;180:787–802.PubMedPubMedCentralCrossRef
195.
Zurück zum Zitat Shintaku J, Peterson JM, Talbert EE, Gu J-M, Ladner KJ, Williams DR, et al. MyoD regulates skeletal muscle oxidative metabolism cooperatively with alternative NF-κ B. Cell Rep. 2016;17:514–26.PubMedPubMedCentralCrossRef Shintaku J, Peterson JM, Talbert EE, Gu J-M, Ladner KJ, Williams DR, et al. MyoD regulates skeletal muscle oxidative metabolism cooperatively with alternative NF-κ B. Cell Rep. 2016;17:514–26.PubMedPubMedCentralCrossRef
196.
Zurück zum Zitat Londhe P, Yu PY, Ijiri Y, Ladner KJ, Fenger JM, London C, et al. Classical NF-κB metabolically reprograms sarcoma cells through regulation of hexokinase 2. Front Oncol. 2018;8:104.PubMedPubMedCentralCrossRef Londhe P, Yu PY, Ijiri Y, Ladner KJ, Fenger JM, London C, et al. Classical NF-κB metabolically reprograms sarcoma cells through regulation of hexokinase 2. Front Oncol. 2018;8:104.PubMedPubMedCentralCrossRef
197.
Zurück zum Zitat Jahnke VE, Sabido O, Defour A, Castells J, Lefai E, Roussel D, et al. Evidence for mitochondrial respiratory deficiency in rat rhabdomyosarcoma cells. PLoS ONE. 2010;5:e8637.PubMedPubMedCentralCrossRef Jahnke VE, Sabido O, Defour A, Castells J, Lefai E, Roussel D, et al. Evidence for mitochondrial respiratory deficiency in rat rhabdomyosarcoma cells. PLoS ONE. 2010;5:e8637.PubMedPubMedCentralCrossRef
198.
Zurück zum Zitat Aspuria P-JP, Lunt SY, Väremo L, Vergnes L, Gozo M, Beach JA, et al. Succinate dehydrogenase inhibition leads to epithelial–mesenchymal transition and reprogrammed carbon metabolism. Cancer Metab. 2014;2:21.PubMedPubMedCentralCrossRef Aspuria P-JP, Lunt SY, Väremo L, Vergnes L, Gozo M, Beach JA, et al. Succinate dehydrogenase inhibition leads to epithelial–mesenchymal transition and reprogrammed carbon metabolism. Cancer Metab. 2014;2:21.PubMedPubMedCentralCrossRef
199.
Zurück zum Zitat Sulkowski PL, Oeck S, Dow J, Economos NG, Mirfakhraie L, Liu Y, et al. Oncometabolites suppress DNA repair by disrupting local chromatin signalling. Nature. 2020;582:586–91.PubMedPubMedCentralCrossRef Sulkowski PL, Oeck S, Dow J, Economos NG, Mirfakhraie L, Liu Y, et al. Oncometabolites suppress DNA repair by disrupting local chromatin signalling. Nature. 2020;582:586–91.PubMedPubMedCentralCrossRef
200.
Zurück zum Zitat Salminen A, Kauppinen A, Kaarniranta K. 2-Oxoglutarate-dependent dioxygenases are sensors of energy metabolism, oxygen availability, and iron homeostasis: potential role in the regulation of aging process. Cell Mol Life Sci. 2015;72:3897–914.PubMedCrossRef Salminen A, Kauppinen A, Kaarniranta K. 2-Oxoglutarate-dependent dioxygenases are sensors of energy metabolism, oxygen availability, and iron homeostasis: potential role in the regulation of aging process. Cell Mol Life Sci. 2015;72:3897–914.PubMedCrossRef
201.
Zurück zum Zitat Pollock RE, Randall RL, O’Sullivan B. Sarcoma oncology: a multidisciplinary approach. New York: PMPH USA; 2019. Pollock RE, Randall RL, O’Sullivan B. Sarcoma oncology: a multidisciplinary approach. New York: PMPH USA; 2019.
202.
Zurück zum Zitat Hoffmann A-C, Danenberg KD, Taubert H, Danenberg PV, Wuerl P. A three-gene signature for outcome in soft tissue sarcoma. Clin Cancer Res. 2009;15:5191–8.PubMedCrossRef Hoffmann A-C, Danenberg KD, Taubert H, Danenberg PV, Wuerl P. A three-gene signature for outcome in soft tissue sarcoma. Clin Cancer Res. 2009;15:5191–8.PubMedCrossRef
203.
Zurück zum Zitat Corless CL, Barnett CM, Heinrich MC. Gastrointestinal stromal tumours: origin and molecular oncology. Nat Rev Cancer. 2011;11:865–78.PubMedCrossRef Corless CL, Barnett CM, Heinrich MC. Gastrointestinal stromal tumours: origin and molecular oncology. Nat Rev Cancer. 2011;11:865–78.PubMedCrossRef
204.
Zurück zum Zitat Sadri N, Zhang P. Hypoxia-inducible factors: mediators of cancer progression; prognostic and therapeutic targets in soft tissue sarcomas. Cancers. 2013;5:320–33.PubMedPubMedCentralCrossRef Sadri N, Zhang P. Hypoxia-inducible factors: mediators of cancer progression; prognostic and therapeutic targets in soft tissue sarcomas. Cancers. 2013;5:320–33.PubMedPubMedCentralCrossRef
205.
Zurück zum Zitat Das B, Tsuchida R, Malkin D, Koren G, Baruchel S, Yeger H. Hypoxia enhances tumor stemness by increasing the invasive and tumorigenic side population fraction. Stem Cells. 2008;26:1818–30.PubMedCrossRef Das B, Tsuchida R, Malkin D, Koren G, Baruchel S, Yeger H. Hypoxia enhances tumor stemness by increasing the invasive and tumorigenic side population fraction. Stem Cells. 2008;26:1818–30.PubMedCrossRef
206.
Zurück zum Zitat Bott AJ, Maimouni S, Zong W-X. The pleiotropic effects of glutamine metabolism in cancer. Cancers (Basel). 2019;11:770.CrossRef Bott AJ, Maimouni S, Zong W-X. The pleiotropic effects of glutamine metabolism in cancer. Cancers (Basel). 2019;11:770.CrossRef
207.
Zurück zum Zitat Jackson M, Serada N, Sheehan M, Srinivasan S, Mason N, Guha M, et al. Mitochondrial genome and functional defects in osteosarcoma are associated with their aggressive phenotype. PLoS ONE. 2018;13:e0209489.PubMedPubMedCentralCrossRef Jackson M, Serada N, Sheehan M, Srinivasan S, Mason N, Guha M, et al. Mitochondrial genome and functional defects in osteosarcoma are associated with their aggressive phenotype. PLoS ONE. 2018;13:e0209489.PubMedPubMedCentralCrossRef
208.
Zurück zum Zitat Srinivasan S, Guha M, Dong DW, Whelan KA, Ruthel G, Uchikado Y, et al. Disruption of cytochrome c oxidase function induces the Warburg effect and metabolic reprogramming. Oncogene. 2016;35:1585–95.PubMedCrossRef Srinivasan S, Guha M, Dong DW, Whelan KA, Ruthel G, Uchikado Y, et al. Disruption of cytochrome c oxidase function induces the Warburg effect and metabolic reprogramming. Oncogene. 2016;35:1585–95.PubMedCrossRef
209.
Zurück zum Zitat Srinivasan S, Guha M, Kashina A, Avadhani NG. Mitochondrial dysfunction and mitochondrial dynamics—the cancer connection. Biochim Biophys Acta (BBA) Bioenerg. 2017;1858:602–14.CrossRef Srinivasan S, Guha M, Kashina A, Avadhani NG. Mitochondrial dysfunction and mitochondrial dynamics—the cancer connection. Biochim Biophys Acta (BBA) Bioenerg. 2017;1858:602–14.CrossRef
210.
Zurück zum Zitat Guha M, Srinivasan S, Ruthel G, Kashina AK, Carstens RP, Mendoza A, et al. Mitochondrial retrograde signaling induces epithelial–mesenchymal transition and generates breast cancer stem cells. Oncogene. 2014;33:5238–50.PubMedCrossRef Guha M, Srinivasan S, Ruthel G, Kashina AK, Carstens RP, Mendoza A, et al. Mitochondrial retrograde signaling induces epithelial–mesenchymal transition and generates breast cancer stem cells. Oncogene. 2014;33:5238–50.PubMedCrossRef
211.
Zurück zum Zitat Yizhak K, Le Dévédec SE, Rogkoti VM, Baenke F, Boer VC, Frezza C, et al. A computational study of the Warburg effect identifies metabolic targets inhibiting cancer migration. Mol Syst Biol. 2014;10:744.PubMedPubMedCentralCrossRef Yizhak K, Le Dévédec SE, Rogkoti VM, Baenke F, Boer VC, Frezza C, et al. A computational study of the Warburg effect identifies metabolic targets inhibiting cancer migration. Mol Syst Biol. 2014;10:744.PubMedPubMedCentralCrossRef
212.
213.
Zurück zum Zitat Kalluri R. The biology and function of fibroblasts in cancer. Nat Rev Cancer. 2016;16:582–98.PubMedCrossRef Kalluri R. The biology and function of fibroblasts in cancer. Nat Rev Cancer. 2016;16:582–98.PubMedCrossRef
214.
Zurück zum Zitat Bittner JG, Wilson M, Shah MB, Albo D, Feig BW, Wang TN. Fibroblast-conditioned media promote human sarcoma cell invasion. Surgery. 2009;145:42–7.PubMedCrossRef Bittner JG, Wilson M, Shah MB, Albo D, Feig BW, Wang TN. Fibroblast-conditioned media promote human sarcoma cell invasion. Surgery. 2009;145:42–7.PubMedCrossRef
215.
Zurück zum Zitat Bonuccelli G, Avnet S, Grisendi G, Salerno M, Granchi D, Dominici M, et al. Role of mesenchymal stem cells in osteosarcoma and metabolic reprogramming of tumor cells. Oncotarget. 2014;5:7575–88.PubMedPubMedCentralCrossRef Bonuccelli G, Avnet S, Grisendi G, Salerno M, Granchi D, Dominici M, et al. Role of mesenchymal stem cells in osteosarcoma and metabolic reprogramming of tumor cells. Oncotarget. 2014;5:7575–88.PubMedPubMedCentralCrossRef
216.
Zurück zum Zitat Bonuccelli G, Tsirigos A, Whitaker-Menezes D, Pavlides S, Pestell RG, Chiavarina B, et al. Ketones and lactate “fuel” tumor growth and metastasis. Cell Cycle. 2014;9:9. Bonuccelli G, Tsirigos A, Whitaker-Menezes D, Pavlides S, Pestell RG, Chiavarina B, et al. Ketones and lactate “fuel” tumor growth and metastasis. Cell Cycle. 2014;9:9.
217.
Zurück zum Zitat Dai L, Qin Z, Defee M, Toole BP, Kirkwood KL, Parsons C. Kaposi sarcoma-associated herpesvirus (KSHV) induces a functional tumor-associated phenotype for oral fibroblasts. Cancer Lett. 2012;318:214–20.PubMedCrossRef Dai L, Qin Z, Defee M, Toole BP, Kirkwood KL, Parsons C. Kaposi sarcoma-associated herpesvirus (KSHV) induces a functional tumor-associated phenotype for oral fibroblasts. Cancer Lett. 2012;318:214–20.PubMedCrossRef
218.
Zurück zum Zitat De Saedeleer CJ, Copetti T, Porporato PE, Verrax J, Feron O, Sonveaux P. Lactate activates HIF-1 in oxidative but not in Warburg-phenotype human tumor cells. PLoS ONE. 2012;7:e46571.PubMedPubMedCentralCrossRef De Saedeleer CJ, Copetti T, Porporato PE, Verrax J, Feron O, Sonveaux P. Lactate activates HIF-1 in oxidative but not in Warburg-phenotype human tumor cells. PLoS ONE. 2012;7:e46571.PubMedPubMedCentralCrossRef
220.
Zurück zum Zitat Goodwin ML, Jin H, Straessler K, Smith-Fry K, Zhu J-F, Monument MJ, et al. Modeling alveolar soft part sarcomagenesis in the mouse: a role for lactate in the tumor microenvironment. Cancer Cell. 2014;26:851–62.PubMedPubMedCentralCrossRef Goodwin ML, Jin H, Straessler K, Smith-Fry K, Zhu J-F, Monument MJ, et al. Modeling alveolar soft part sarcomagenesis in the mouse: a role for lactate in the tumor microenvironment. Cancer Cell. 2014;26:851–62.PubMedPubMedCentralCrossRef
221.
Zurück zum Zitat Porporato PE. Mitochondrial metabolism and cancer. Cell Res. 2018;28:16.CrossRef Porporato PE. Mitochondrial metabolism and cancer. Cell Res. 2018;28:16.CrossRef
222.
Zurück zum Zitat Danhier P, Bański P, Payen VL, Grasso D, Ippolito L, Sonveaux P, et al. Cancer metabolism in space and time: beyond the Warburg effect. Biochim Biophys Acta (BBA) Bioenerg. 2017;1858:556–72.CrossRef Danhier P, Bański P, Payen VL, Grasso D, Ippolito L, Sonveaux P, et al. Cancer metabolism in space and time: beyond the Warburg effect. Biochim Biophys Acta (BBA) Bioenerg. 2017;1858:556–72.CrossRef
223.
Zurück zum Zitat Porporato PE. Metabolic changes associated with tumor metastasis, part 2: mitochondria, lipid and amino acid metabolism. Cell Mol Life Sci. 2016;73:1349–63.PubMedCrossRef Porporato PE. Metabolic changes associated with tumor metastasis, part 2: mitochondria, lipid and amino acid metabolism. Cell Mol Life Sci. 2016;73:1349–63.PubMedCrossRef
224.
Zurück zum Zitat Harati K, Daigeler A, Hirsch T, Jacobsen F, Behr B, Wallner C, et al. Tumor-associated fibroblasts promote the proliferation and decrease the doxorubicin sensitivity of liposarcoma cells. Int J Mol Med. 2016;37:1535–41.PubMedPubMedCentralCrossRef Harati K, Daigeler A, Hirsch T, Jacobsen F, Behr B, Wallner C, et al. Tumor-associated fibroblasts promote the proliferation and decrease the doxorubicin sensitivity of liposarcoma cells. Int J Mol Med. 2016;37:1535–41.PubMedPubMedCentralCrossRef
225.
Zurück zum Zitat Bellairs R, Van Peteghem M-C. Gastrulation: is it analogous to malignant invasion. Am Zool. 1984;24:563–70.CrossRef Bellairs R, Van Peteghem M-C. Gastrulation: is it analogous to malignant invasion. Am Zool. 1984;24:563–70.CrossRef
226.
Zurück zum Zitat Sannino G, Marchetto A, Kirchner T, Grünewald TGP. Epithelial-to-mesenchymal and mesenchymal-to-epithelial transition in mesenchymal tumors: a paradox in sarcomas? Cancer Res. 2017;77:4556–61.PubMedCrossRef Sannino G, Marchetto A, Kirchner T, Grünewald TGP. Epithelial-to-mesenchymal and mesenchymal-to-epithelial transition in mesenchymal tumors: a paradox in sarcomas? Cancer Res. 2017;77:4556–61.PubMedCrossRef
227.
Zurück zum Zitat Chaklader M, Pan A, Law A, Chattopadhayay S, Chatterjee R, Law S. Differential remodeling of cadherins and intermediate cytoskeletal filaments influence microenvironment of solid and ascitic sarcoma. Mol Cell Biochem. 2013;382:293–306.PubMedCrossRef Chaklader M, Pan A, Law A, Chattopadhayay S, Chatterjee R, Law S. Differential remodeling of cadherins and intermediate cytoskeletal filaments influence microenvironment of solid and ascitic sarcoma. Mol Cell Biochem. 2013;382:293–306.PubMedCrossRef
228.
Zurück zum Zitat Kahlert UD, Joseph JV, Kruyt FAE. EMT- and MET-related processes in nonepithelial tumors: importance for disease progression, prognosis, and therapeutic opportunities. Mol Oncol. 2017;11:860–77.PubMedPubMedCentralCrossRef Kahlert UD, Joseph JV, Kruyt FAE. EMT- and MET-related processes in nonepithelial tumors: importance for disease progression, prognosis, and therapeutic opportunities. Mol Oncol. 2017;11:860–77.PubMedPubMedCentralCrossRef
229.
Zurück zum Zitat Tian W, Wang G, Yang J, Pan Y, Ma Y. Prognostic role of E-cadherin and Vimentin expression in various subtypes of soft tissue leiomyosarcomas. Med Oncol. 2013;30:401.PubMedCrossRef Tian W, Wang G, Yang J, Pan Y, Ma Y. Prognostic role of E-cadherin and Vimentin expression in various subtypes of soft tissue leiomyosarcomas. Med Oncol. 2013;30:401.PubMedCrossRef
230.
Zurück zum Zitat Saito T. The SYT-SSX fusion protein and histological epithelial differentiation in synovial sarcoma: relationship with extracellular matrix remodeling. Int J Clin Exp Pathol. 2013;6:2272.PubMedPubMedCentral Saito T. The SYT-SSX fusion protein and histological epithelial differentiation in synovial sarcoma: relationship with extracellular matrix remodeling. Int J Clin Exp Pathol. 2013;6:2272.PubMedPubMedCentral
231.
Zurück zum Zitat Thuault S, Hayashi S, Lagirand-Cantaloube J, Plutoni C, Comunale F, Delattre O, et al. P-cadherin is a direct PAX3–FOXO1A target involved in alveolar rhabdomyosarcoma aggressiveness. Oncogene. 2013;32:1876–87.PubMedCrossRef Thuault S, Hayashi S, Lagirand-Cantaloube J, Plutoni C, Comunale F, Delattre O, et al. P-cadherin is a direct PAX3–FOXO1A target involved in alveolar rhabdomyosarcoma aggressiveness. Oncogene. 2013;32:1876–87.PubMedCrossRef
232.
Zurück zum Zitat Hua W, ten Dijke P, Kostidis S, Giera M, Hornsveld M. TGFβ-induced metabolic reprogramming during epithelial-to-mesenchymal transition in cancer. Cell Mol Life Sci. 2020;77:2103–23.PubMedCrossRef Hua W, ten Dijke P, Kostidis S, Giera M, Hornsveld M. TGFβ-induced metabolic reprogramming during epithelial-to-mesenchymal transition in cancer. Cell Mol Life Sci. 2020;77:2103–23.PubMedCrossRef
233.
234.
Zurück zum Zitat El-Naggar AM, Veinotte CJ, Cheng H, Grunewald TGP, Negri GL, Somasekharan SP, et al. Translational activation of HIF1α by YB-1 promotes sarcoma metastasis. Cancer Cell. 2015;27:682–97.PubMedCrossRef El-Naggar AM, Veinotte CJ, Cheng H, Grunewald TGP, Negri GL, Somasekharan SP, et al. Translational activation of HIF1α by YB-1 promotes sarcoma metastasis. Cancer Cell. 2015;27:682–97.PubMedCrossRef
235.
236.
Zurück zum Zitat Eisinger-Mathason TSK, Zhang M, Qiu Q, Skuli N, Nakazawa MS, Karakasheva T, et al. Hypoxia-dependent modification of collagen networks promotes sarcoma metastasis. Cancer Discov. 2013;3:1190–205.PubMedCrossRef Eisinger-Mathason TSK, Zhang M, Qiu Q, Skuli N, Nakazawa MS, Karakasheva T, et al. Hypoxia-dependent modification of collagen networks promotes sarcoma metastasis. Cancer Discov. 2013;3:1190–205.PubMedCrossRef
237.
Zurück zum Zitat Petitprez F, Meylan M, de Reyniès A, Sautès-Fridman C, Fridman WH. The tumor microenvironment in the response to immune checkpoint blockade therapies. Front Immunol. 2020;11:784.PubMedPubMedCentralCrossRef Petitprez F, Meylan M, de Reyniès A, Sautès-Fridman C, Fridman WH. The tumor microenvironment in the response to immune checkpoint blockade therapies. Front Immunol. 2020;11:784.PubMedPubMedCentralCrossRef
238.
239.
Zurück zum Zitat Chalmers ZR, Connelly CF, Fabrizio D, Gay L, Ali SM, Ennis R, et al. Analysis of 100,000 human cancer genomes reveals the landscape of tumor mutational burden. Genome Med. 2017;9:34.PubMedPubMedCentralCrossRef Chalmers ZR, Connelly CF, Fabrizio D, Gay L, Ali SM, Ennis R, et al. Analysis of 100,000 human cancer genomes reveals the landscape of tumor mutational burden. Genome Med. 2017;9:34.PubMedPubMedCentralCrossRef
240.
241.
Zurück zum Zitat Fletcher CDM. The evolving classification of soft tissue tumours—an update based on the new 2013 WHO classification. Histopathology. 2014;64:2–11.PubMedCrossRef Fletcher CDM. The evolving classification of soft tissue tumours—an update based on the new 2013 WHO classification. Histopathology. 2014;64:2–11.PubMedCrossRef
242.
Zurück zum Zitat Cohen JE, Eleyan F, Zick A, Peretz T, Katz D. Intratumoral immune-biomarkers and mismatch repair status in leiyomyosarcoma -potential predictive markers for adjuvant treatment: a pilot study. Oncotarget. 2018;9:30847–54.PubMedPubMedCentralCrossRef Cohen JE, Eleyan F, Zick A, Peretz T, Katz D. Intratumoral immune-biomarkers and mismatch repair status in leiyomyosarcoma -potential predictive markers for adjuvant treatment: a pilot study. Oncotarget. 2018;9:30847–54.PubMedPubMedCentralCrossRef
243.
Zurück zum Zitat D’Angelo SP, Tap WD, Schwartz GK, Carvajal RD. Sarcoma immunotherapy: past approaches and future directions. Sarcoma. 2014;2014:1–13.CrossRef D’Angelo SP, Tap WD, Schwartz GK, Carvajal RD. Sarcoma immunotherapy: past approaches and future directions. Sarcoma. 2014;2014:1–13.CrossRef
244.
Zurück zum Zitat van Erp AEM, Versleijen-Jonkers YMH, Hillebrandt-Roeffen MHS, van Houdt L, Gorris MAJ, van Dam LS, et al. Expression and clinical association of programmed cell death-1, programmed death-ligand-1 and CD8+ lymphocytes in primary sarcomas is subtype dependent. Oncotarget. 2017;8:71371–84.PubMedPubMedCentralCrossRef van Erp AEM, Versleijen-Jonkers YMH, Hillebrandt-Roeffen MHS, van Houdt L, Gorris MAJ, van Dam LS, et al. Expression and clinical association of programmed cell death-1, programmed death-ligand-1 and CD8+ lymphocytes in primary sarcomas is subtype dependent. Oncotarget. 2017;8:71371–84.PubMedPubMedCentralCrossRef
245.
Zurück zum Zitat Wedekind MF, Wagner LM, Cripe TP. Immunotherapy for osteosarcoma: where do we go from here? Pediatr Blood Cancer. 2018;65:e27227.PubMedCrossRef Wedekind MF, Wagner LM, Cripe TP. Immunotherapy for osteosarcoma: where do we go from here? Pediatr Blood Cancer. 2018;65:e27227.PubMedCrossRef
246.
Zurück zum Zitat Dancsok AR, Setsu N, Gao D, Blay J-Y, Thomas D, Maki RG, et al. Expression of lymphocyte immunoregulatory biomarkers in bone and soft-tissue sarcomas. Mod Pathol. 2019;32:1772–85.PubMedCrossRef Dancsok AR, Setsu N, Gao D, Blay J-Y, Thomas D, Maki RG, et al. Expression of lymphocyte immunoregulatory biomarkers in bone and soft-tissue sarcomas. Mod Pathol. 2019;32:1772–85.PubMedCrossRef
247.
Zurück zum Zitat Feng X, Pleasance E, Zhao EY, Ng T, Grewal JK, Mohammad N, et al. Therapeutic implication of genomic landscape of adult metastatic sarcoma. JCO Precis Oncol. 2019;3:1–25.PubMed Feng X, Pleasance E, Zhao EY, Ng T, Grewal JK, Mohammad N, et al. Therapeutic implication of genomic landscape of adult metastatic sarcoma. JCO Precis Oncol. 2019;3:1–25.PubMed
248.
Zurück zum Zitat Keung EZ, Tsai J-W, Ali AM, Cormier JN, Bishop AJ, Guadagnolo BA, et al. Analysis of the immune infiltrate in undifferentiated pleomorphic sarcoma of the extremity and trunk in response to radiotherapy: rationale for combination neoadjuvant immune checkpoint inhibition and radiotherapy. Oncoimmunology. 2017;7:e1385689.PubMedPubMedCentralCrossRef Keung EZ, Tsai J-W, Ali AM, Cormier JN, Bishop AJ, Guadagnolo BA, et al. Analysis of the immune infiltrate in undifferentiated pleomorphic sarcoma of the extremity and trunk in response to radiotherapy: rationale for combination neoadjuvant immune checkpoint inhibition and radiotherapy. Oncoimmunology. 2017;7:e1385689.PubMedPubMedCentralCrossRef
249.
Zurück zum Zitat Sautès-Fridman C, Petitprez F, Calderaro J, Fridman WH. Tertiary lymphoid structures in the era of cancer immunotherapy. Nat Rev Cancer. 2019;19:307–25.PubMedCrossRef Sautès-Fridman C, Petitprez F, Calderaro J, Fridman WH. Tertiary lymphoid structures in the era of cancer immunotherapy. Nat Rev Cancer. 2019;19:307–25.PubMedCrossRef
250.
Zurück zum Zitat Varn FS, Wang Y, Mullins DW, Fiering S, Cheng C. Systematic pan-cancer analysis reveals immune cell interactions in the tumor microenvironment. Cancer Res. 2017;77:1271–82.PubMedPubMedCentralCrossRef Varn FS, Wang Y, Mullins DW, Fiering S, Cheng C. Systematic pan-cancer analysis reveals immune cell interactions in the tumor microenvironment. Cancer Res. 2017;77:1271–82.PubMedPubMedCentralCrossRef
252.
Zurück zum Zitat Hayward SL, Scharer CD, Cartwright EK, Takamura S, Li Z-RT, Boss JM, et al. Environmental cues regulate epigenetic reprogramming of airway-resident memory CD8+ T cells. Nat Immunol. 2020;21:309–20.PubMedPubMedCentralCrossRef Hayward SL, Scharer CD, Cartwright EK, Takamura S, Li Z-RT, Boss JM, et al. Environmental cues regulate epigenetic reprogramming of airway-resident memory CD8+ T cells. Nat Immunol. 2020;21:309–20.PubMedPubMedCentralCrossRef
253.
Zurück zum Zitat Yan D, Adeshakin AO, Xu M, Afolabi LO, Zhang G, Chen YH, et al. Lipid metabolic pathways confer the immunosuppressive function of myeloid-derived suppressor cells in tumor. Front Immunol. 2019;10:1399.PubMedPubMedCentralCrossRef Yan D, Adeshakin AO, Xu M, Afolabi LO, Zhang G, Chen YH, et al. Lipid metabolic pathways confer the immunosuppressive function of myeloid-derived suppressor cells in tumor. Front Immunol. 2019;10:1399.PubMedPubMedCentralCrossRef
254.
Zurück zum Zitat Weinberg SE, Singer BD, Steinert EM, Martinez CA, Mehta MM, Martínez-Reyes I, et al. Mitochondrial complex III is essential for suppressive function of regulatory T cells. Nature. 2019;565:495–9.PubMedPubMedCentralCrossRef Weinberg SE, Singer BD, Steinert EM, Martinez CA, Mehta MM, Martínez-Reyes I, et al. Mitochondrial complex III is essential for suppressive function of regulatory T cells. Nature. 2019;565:495–9.PubMedPubMedCentralCrossRef
255.
Zurück zum Zitat Kumagai S, Togashi Y, Sakai C, Kawazoe A, Kawazu M, Ueno T, et al. An oncogenic alteration creates a microenvironment that promotes tumor progression by conferring a metabolic advantage to regulatory T cells. Immunity. 2020;53:187-203.e8.PubMedCrossRef Kumagai S, Togashi Y, Sakai C, Kawazoe A, Kawazu M, Ueno T, et al. An oncogenic alteration creates a microenvironment that promotes tumor progression by conferring a metabolic advantage to regulatory T cells. Immunity. 2020;53:187-203.e8.PubMedCrossRef
256.
257.
Zurück zum Zitat Naquet P, Giessner C, Galland F. Metabolic adaptation of tissues to stress releases metabolites influencing innate immunity. Curr Opin Immunol. 2016;38:30–8.PubMedCrossRef Naquet P, Giessner C, Galland F. Metabolic adaptation of tissues to stress releases metabolites influencing innate immunity. Curr Opin Immunol. 2016;38:30–8.PubMedCrossRef
258.
Zurück zum Zitat Wang H, Franco F, Tsui Y-C, Xie X, Trefny MP, Zappasodi R, et al. CD36-mediated metabolic adaptation supports regulatory T cell survival and function in tumors. Nat Immunol. 2020;21:23.CrossRef Wang H, Franco F, Tsui Y-C, Xie X, Trefny MP, Zappasodi R, et al. CD36-mediated metabolic adaptation supports regulatory T cell survival and function in tumors. Nat Immunol. 2020;21:23.CrossRef
Metadaten
Titel
Metabolic landscapes in sarcomas
verfasst von
Richard Miallot
Franck Galland
Virginie Millet
Jean-Yves Blay
Philippe Naquet
Publikationsdatum
01.12.2021
Verlag
BioMed Central
Erschienen in
Journal of Hematology & Oncology / Ausgabe 1/2021
Elektronische ISSN: 1756-8722
DOI
https://doi.org/10.1186/s13045-021-01125-y

Weitere Artikel der Ausgabe 1/2021

Journal of Hematology & Oncology 1/2021 Zur Ausgabe

Adjuvante Immuntherapie verlängert Leben bei RCC

25.04.2024 Nierenkarzinom Nachrichten

Nun gibt es auch Resultate zum Gesamtüberleben: Eine adjuvante Pembrolizumab-Therapie konnte in einer Phase-3-Studie das Leben von Menschen mit Nierenzellkarzinom deutlich verlängern. Die Sterberate war im Vergleich zu Placebo um 38% geringer.

Alectinib verbessert krankheitsfreies Überleben bei ALK-positivem NSCLC

25.04.2024 NSCLC Nachrichten

Das Risiko für Rezidiv oder Tod von Patienten und Patientinnen mit reseziertem ALK-positivem NSCLC ist unter einer adjuvanten Therapie mit dem Tyrosinkinase-Inhibitor Alectinib signifikant geringer als unter platinbasierter Chemotherapie.

Bei Senioren mit Prostatakarzinom auf Anämie achten!

24.04.2024 DGIM 2024 Nachrichten

Patienten, die zur Behandlung ihres Prostatakarzinoms eine Androgendeprivationstherapie erhalten, entwickeln nicht selten eine Anämie. Wer ältere Patienten internistisch mitbetreut, sollte auf diese Nebenwirkung achten.

ICI-Therapie in der Schwangerschaft wird gut toleriert

Müssen sich Schwangere einer Krebstherapie unterziehen, rufen Immuncheckpointinhibitoren offenbar nicht mehr unerwünschte Wirkungen hervor als andere Mittel gegen Krebs.

Update Onkologie

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