Next Article in Journal
Common Kinetic Mechanism of Abasic Site Recognition by Structurally Different Apurinic/Apyrimidinic Endonucleases
Next Article in Special Issue
PTP61F Mediates Cell Competition and Mitigates Tumorigenesis
Previous Article in Journal
Vitamin D Reverses Disruption of Gut Epithelial Barrier Function Caused by Campylobacter jejuni
Previous Article in Special Issue
Drosophila Larval Models of Invasive Tumorigenesis for In Vivo Studies on Tumour/Peripheral Host Tissue Interactions during Cancer Cachexia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

RasV12; scrib−/− Tumors: A Cooperative Oncogenesis Model Fueled by Tumor/Host Interactions

by
Caroline Dillard
1,2,*,
José Gerardo Teles Reis
1,2 and
Tor Erik Rusten
1,2,*
1
Centre for Cancer Cell Reprogramming, Faculty of Medicine, Institute of Clinical Medicine, University of Oslo, 0372 Oslo, Norway
2
Department of Molecular Cell Biology, Institute for Cancer Research, Oslo University Hospital, Montebello, 0379 Oslo, Norway
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(16), 8873; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22168873
Submission received: 9 July 2021 / Revised: 12 August 2021 / Accepted: 12 August 2021 / Published: 18 August 2021
(This article belongs to the Special Issue Basic and Translational Models of Cooperative Oncogenesis 2.0)

Abstract

:
The phenomenon of how oncogenes and tumor-suppressor mutations can synergize to promote tumor fitness and cancer progression can be studied in relatively simple animal model systems such as Drosophila melanogaster. Almost two decades after the landmark discovery of cooperative oncogenesis between oncogenic RasV12 and the loss of the tumor suppressor scribble in flies, this and other tumor models have provided new concepts and findings in cancer biology that has remarkable parallels and relevance to human cancer. Here we review findings using the RasV12; scrib−/− tumor model and how it has contributed to our understanding of how these initial simple genetic insults cooperate within the tumor cell to set in motion the malignant transformation program leading to tumor growth through cell growth, cell survival and proliferation, dismantling of cell–cell interactions, degradation of basement membrane and spreading to other organs. Recent findings have demonstrated that cooperativity goes beyond cell intrinsic mechanisms as the tumor interacts with the immediate cells of the microenvironment, the immune system and systemic organs to eventually facilitate malignant progression.

1. Introduction

Mutations in oncogenes and tumor suppressors can synergize over time in order to optimize the tumor fitness within the host, a phenomenon known as cooperative oncogenesis. For example, cooperation between Kras and Myc has been shown to enhance tumorigenesis and weaken immunosurveillance [1]. Pharmacological inhibition of the cooperation between Kras and mutated p53 has proven efficient in decreasing metastasis [2]. Although these findings highlight the importance of genetic cooperation for malignant transformation, the precise mechanisms operating during cooperative oncogenesis remain unclear.
To investigate unconventional synergizing networks and their physiological outcomes, in vivo models are essential. Because of its short life cycle, the tremendous collection of genetic tools available, and the remarkable conservation of most signaling pathways, Drosophila is a relevant model for studying oncogenesis. During the last decades, cancer research using Drosophila has dramatically increased [3]. Drosophila cancer models as diverse as gut, brain, hematopoietic, and carcinoma models have been developed and exhibit typical hallmarks of cancer [4]. Although some aspects of cancer biology are difficult to model in flies, such as angiogenesis or immunosurveillance, genetic and chemical screens using Drosophila have brought new insights on fundamental cancer biology and the identification of therapeutic targets [3,5].
Imaginal discs from the Drosophila larva constitute a site of choice for investigating the biology of carcinoma—the most common cancer type, which originates from epithelial cells (Figure 1A).
Imaginal discs are simple epithelial layers set aside during embryogenesis. They undergo controlled growth, patterning, differentiation, and morphogenesis during larval and pupal stages to make specific adult organ structures. Due to evolution, the molecular language that controls these processes is conserved among metazoa, from fly to man. It is the regulated growth, patterning, differentiation, programmed cell death, and tissue structure that fall apart during tumorigenesis. This deconstruction can be studied in flies. More importantly, a vast number of studies in the wing and eye-antennal discs have revealed several sets of cooperating genes that control tumor formation, growth and survival (reviewed in [6]).
Among these cancer models, the cooperation between the Drosophila KRAS homolog oncogene, RasV12 and the loss of the cell polarity tumor suppressor scribble (referred to hereafter as “RasV12; scrib−/−”) has been the most extensively investigated. This model recapitulates essential aspects of carcinoma development: loss of epithelial structure and differentiation, sustained proliferation, cytoskeletal remodeling, resistance to cell death, basement membrane degradation, and invasion [7,8,9,10,11]. The study of this specific cooperation has been particularly fruitful, providing novel concepts and mechanisms that go beyond intracellular cooperation.
In this review, we summarize the contribution of the RasV12; scrib−/− model to the understanding of cooperative oncogenesis and cancer biology. We start with the first observations of the cooperation between RasV12 oncogene expression and the loss of cell polarity. We then review how intrinsic cues contribute to the growth of malignant RasV12; scrib−/− tumors. Subsequently, we highlight the importance of local interactions between the tumor and its microenvironment for regulating its fate and progression. Finally, we consider the effect of long-range tumor–host interactions for sustaining tumor growth and transformation.

2. Establishment of the RasV12; scrib−/− Cooperative Model

The RasV12; scrib−/− tumor model is a good illustration of the concept of cooperative oncogenesis. Alone, each insult (loss of cell polarity or RasV12 overexpression) gives rise to slow-growing non-invasive benign tumors. However, together, they trigger the growth of massive invasive malignant tumors that ultimately kill the host. In this part, we describe the first observations that led to the establishment of this cooperative oncogenic model and their relevance to mammalian cancers.

2.1. Loss of the Cell Polarity Gene Scribble Triggers the Formation of Benign Tumors

Downregulation and mislocalization of cell polarity proteins is typically observed in many human tumors (reviewed in [12]). It often appears during the early stages of malignancy [13,14] and is associated with aggressive cancers [15,16,17,18,19,20,21]. Conversely, polarity gene amplification has been reported in several human cancers. These opposite observations highlight the need for a better understanding of the function of cell polarity regulators in cancer and the characterization of their downstream effectors in different contexts [22].
Studies in Drosophila have allowed the discovery of the Scribble polarity tricomplex, which is composed of the LAP scaffolding protein Scribble (Scrib), the MAGUK protein Disc large (Dlg), and the WD-repeat protein Lethal giant larvae (Lgl). In Drosophila and mammals, the three proteins are located at the basolateral side of epithelial cells (Reviewed in [23]). Although physical interactions between these proteins are not yet fully understood, genetic interactions demonstrated that they function together to establish and maintain apicobasal polarity [24]. Apicobasal polarity is essential to define specific membrane compartments which restrict local signal transduction. In this way, individual epithelial cells can integrate cues from their neighboring cells within the epithelium and adjust their proliferation, growth, survival, metabolism, and motility to ensure proper tissue homeostasis (reviewed in [12]). The human orthologs of Scrib, Dlg, and Lgl play similar roles in mammals, and they can substitute for their fly counterparts and rescue loss of cell polarity defects in Drosophila [25,26,27,28,29]. However, their precise modes of action seem more complex, probably due to both redundancies between paralogs and tissue specificities (reviewed in [30]). Interestingly, upon infection, the Human Papilloma Virus targets Scrib and Dlg for degradation, stressing their potential role in tumor formation [31].
Tumor formation can be modeled in Drosophila eye discs using the MARCM technique (Mosaic Analysis with a Repressible Cell Marker) [32]. With this genetic trick, randomly generated individual cells undergo loss of heterozygosity upon mitotic recombination. As a consequence, they can inherit two identical copies of a loss-of-function allele for a tumor suppressor. At the same time, the MARCM technique allows to label the cells with GFP and optionally to drive the expression of any gene of choice, such as an oncogene. Thus, both the loss of a tumor suppressor and the gain of function of an oncogene can be modeled from a single tumor-initiating cell, mimicking the early random genetic events that initiate tumorigenesis in mammals.
MARCM scrib−/− cells form small tumors within the eye disc (Figure 1B). They are characterized by a loss of cell polarity, cell morphology alterations (small round cells, multilayered clones, downregulation of DE-Cadherin), overproliferation involving CyclinE upregulation, and resistance to differentiation (Brumby and Richardson, 2003). These clones, nevertheless, remain small due to their active elimination through apoptosis. As a result, the adult flies harbor a smaller but functional eye. Interestingly, cell polarity mutant clones also display some preinvasive traits. However, despite enrichment for the Matrix Metalloproteinase 1 (MMP1) and subsequent basement membrane degradation, scrib−/− cells are unable to invade other tissues [7,10,11,33,34].
These observations demonstrate that the loss of scrib is not sufficient to drive malignancy in the Drosophila imaginal discs on its own.

2.2. The Expression of the Oncogene RasV12 Alone Is Not Sufficient to Create Neoplastic, Invasive Tumors

Ras is a membrane-associated guanine nucleotide-binding protein. Upon activation by different Receptor Tyrosine Kinase (RTK) growth factor receptors, Ras changes its conformation from a GDP-bound inactive form to an active GTP-bound form that relays the signal to diverse downstream effectors; the MAPK pathway, the PI3K-AKT-TOR pathway, and the Rac-Rho pathway (Reviewed in [35]).
In human cancers, oncogenic altered forms of the proteins N-Ras, H-Ras, and K-Ras aberrantly activate the downstream targets of Ras, leading to uncontrolled proliferation, growth, metabolism, survival, and migration [35]. Mutated Ras proteins are found in 20–30% of human tumors. Significantly, they are often associated with mutations in other genes (such as Myc, tp53, SMAD4), suggesting that mutated Ras alone might not be able to support malignant transformation fully [1,2]. Among these Ras oncogenic versions, the K-Ras protein in which the glycine at codon 12 is mutated into a valine or a serine remains in the GTP-bound constitutive active form. K-Ras is associated with the worst prognosis in many cancer types [4,35,36,37,38].
In Drosophila, there is only one homolog for N-Ras, H-Ras, and K-Ras, called Ras85D. Similar to its mammalian counterpart, wild type Ras85D is activated by many RTKs and regulates growth [39,40,41,42,43], cell identity [39,41,44,45] and survival [42,46] mainly through the Raf/MAPK pathway [47]. However, Ras85D does not seem to promote proliferation at physiological levels and is likely uncoupled from the PI3K/AKT pathway [40,48]. Similar to K-Ras Codon 12 -mutated proteins, the engineered Drosophila RasV12 allele where valine is substituted with glycine at position 12 produces a constitutively active form of Ras85D [43].
RasV12 MARCM clones in the eye antennal disc grow moderately and overproliferate to form hyperplastic tumors [11] (Figure 1C). In these tumors, RasV12 increases dMyc levels through Raf/MAPK signaling. Concomitantly and contrary to wild type Ras85D, RasV12 ectopically activates PI3K/AKT signaling, which might explain why only RasV12 and not Ras85D expression stimulates proliferation [48]. RasV12 tumor-bearing larvae can initiate metamorphosis on time. However, most of them die before reaching adulthood, probably because of developmental defects [49]. Interestingly, RasV12 clones trigger non-autonomous cell death in the surrounding wild type (wt) cells indicating that there is a sensing mechanism that mediates the detection of transformed RasV12 cells by wt cells [50] (see the section on short-range interaction and cell competition). The few adult escapers display deformed eyes and head structures with necrotic patches [11]. Of note, in the wing disc, RasV12 cells seem to exhibit some invasive properties as they can migrate through the disc basement membrane. Nevertheless, metalloproteinase production and basement membrane degradation has not yet been demonstrated. RasV12 transformed cells can also be extruded on the apical side [51]. However, secondary tumor formation has never been observed in this context, indicating that although RasV12 cells can manage to escape their primary site, they cannot invade other organs and develop into secondary tumors [10,11,52]. Thus, RasV12 clones in the Drosophila imaginal discs form non-invasive benign tumors.

2.3. RasV12 Cooperates with the Loss of Cell Polarity to Drive the Formation of Aggressive Malignant Tumors

Malignant tumors are characterized by uncontrolled growth and the ability to invade and destroy other tissues. In the Drosophila eye-antennal discs, the concomitant loss of cell polarity (scrib−/−) and RasV12 overexpression lead to the formation of aggressive tumors (Figure 1D).
RasV12; scrib−/− tumor-bearing larvae display a prolonged larval stage. Instead of pupariating and undergoing metamorphosis, they extend their larval life and become giant tumor-bearing larvae. During this time, RasV12; scrib−/− tumors grow exponentially, invade and completely cover the adjacent larval central nervous system and leg imaginal discs [7,8,10,11,34,52,53,54,55,56].
Transformed RasV12; scrib−/− cells weaken cell–cell adhesion through the downregulation of DE-cadherin [11] and display migratory characteristics such as F-actin enrichment and filopodia formation [7,10,11].
Simultaneously, inside RasV12; scrib−/− tumors, a subset of cells secrete the metalloproteinase MMP1, which triggers the degradation of the basement membrane [7,8,9,10,11] and allows cells to escape and invade other organs.
In tumor-bearing larvae, RasV12; scrib−/− cells are found in the ventral nerve cord, 1st and 2nd leg discs, trachea, mouth hook, salivary gland, gut, and fat body. They are highly motile, so that some cells are also able to roam freely in the hemolymph [7,10,11,53].
Strikingly, when given more time to grow thanks to their transplantation into the abdomen of adult flies, RasV12; scrib−/− tumors overproliferate and form a big mass that kills the host by approximately ten days. This contrasts to RasV12 or scrib−/− singly mutated tumors. Transplanted RasV12; scrib−/− cells retain their migratory characteristic and are found inside the gut wall and the ovarian follicle, a location only reached after several days [11,53].
Taken together, these observations demonstrate how the cooperation between RasV12 overexpression and loss of scrib is essential for the development of malignant tumors that proliferate indefinitely and invade other organs (Figure 1).
The study of the genetic interactions between the two drivers of this model for cooperative oncogenesis has rapidly allowed the identification of key downstream signaling pathways and new concepts, which together explain malignancy.

3. Intrinsic Effectors of Malignant Transformation

In this part, we review the intrinsic cues that control RasV12; scrib−/− cell transformation. We first describe the central yet ambivalent role of JNK signaling in the development of benign vs. malignant tumors. We then highlight the function of the JAK-STAT pathway in this context. Finally, we focus on the role of Salvador-Warts-Hippo signaling.

3.1. The JNK Pathway, a Coordinator of RasV12; scrib−/− Malignant Transformation

3.1.1. The Ambivalent Role of JNK Signaling in Benign Tumors vs. Malignant Tumors

High levels of cJun N-terminal Kinase (JNK) signaling have been observed inside scrib−/− clones [33,57] and RasV12; scrib−/− clones [7], whereas only basal levels of JNK signaling have been reported in RasV12 clones [55,58].
Intriguingly, JNK signaling inhibition in scrib−/− clones leads to the formation of big lethal tumors [10,33,34], whereas in RasV12; scrib−/− clones, JNK repression gives rise to small non-invasive tumors [9,10].
Thus, JNK signaling has an anti- and a pro-tumorigenic function in scrib−/− and RasV12; scrib−/− clones, respectively.

3.1.2. Drosophila JNK Signaling Overview

The well-conserved JNK pathway integrates a huge diversity of inputs and promotes various processes such as apoptosis, growth, differentiation inhibition, or migration (reviewed in [59]).
In Drosophila, JNK signaling activation results from diverse cues such as oxidative stress, caspases, or binding of the Drosophila Tumor Necrosis Factor (TNF) ligand Eiger [33,60,61] to its TNF receptors (TNFR) Wengen (Wgn) [62] or Grindewald (Grnd) [63].
In the Drosophila eye disc, Eiger overexpression triggers JNK-dependent apoptosis leading to a smaller eye phenotype. This model allowed the characterization of the Eiger-downstream JNK components that are competent in this tissue. Eiger is able to bind to both TNFRs Wgn and Grnd. Whether these receptors possess distinct functions within the pathway is still not clear. However, it has recently been shown that the different binding affinities of Eiger to its receptors govern their internalization trajectory within the cell, which would promote different JNK signaling outcomes [64]. This observation suggests that Wgn and Grnd might have distinct roles in the eye disc. Nevertheless, Egr-induced cell death in the eye disc seems to require both Wgn [62,65] and Grnd [63]. Upon binding to Eiger, the TNFRs activate a kinase cascade including the JNK Kinase Kinase Kinase (JNKKKK) Misshapen (Msn), the JNK Kinase Kinase (JNKKK) Tak1, and the JNK Kinase (JNKK) Hep leading to the final activation of the single JNK, Basket (Bsk). This has long been considered the canonical JNK pathway in the eye disc (Reviewed in [66]). Nonetheless, one might argue that ectopic Eiger expression in the eye disc might lead to aberrant JNK signaling, which does not represent the complete JNK pathway network competent in the eye disc. Notably, more studies in other tissues and contexts have further expanded the list of JNK signaling upstream components in Drosophila at every level of the pathway [63,67]).
Once activated, Bsk phosphorylates several transcription factors (TFs). The most famous ones are the TFs Kayak (Kay aka Fos) and Jun-related antigen (Jra), which heterodimerize to form the so-called AP-1 complex. Transcriptional targets include the proapoptotic genes reaper and hid, the negative JNK regulator puckered (puc), the JNKKKK msn, the matrix metalloproteinase MMP1, the integrin-associated gene paxillin, the JAK-STAT ligands upd1, upd2 and upd3 and the Wnt signaling ligand wingless (wg) (reviewed in [59].

3.1.3. JNK Signaling Functions in scrib−/− Versus RasV12; scrib−/− Tumors

Given the antagonist observations upon JNK signaling inhibition in scrib−/− versus RasV12; scrib−/− clones, many studies have contributed to deciphering the mechanisms that control JNK pro- and anti-tumorigenic functions.
In scrib−/− clones, JNK is active in most cells and induces apoptosis [10,33,34] most probably through the upregulation of the apoptosis-triggering genes reaper and hid [60,68]. Interestingly, in these dying tumor cells, high levels of endocytosis are observed, which cause the relocalization of Eiger/Wengen complexes from the plasma membrane to endosomal vesicles. This step is required for JNK activation and induction of apoptosis [33]. Simultaneously, JNK promotes the expression of the integrin-associated protein Paxillin and the expression of MMP1, leading to basement membrane degradation [7,33,34,69]. However, scrib−/− cells most likely die before acquiring invasive properties, as no migrative behavior has been observed in these cells. JNK activation inside scrib−/− cells also controls the inhibition of photoneuron differentiation [33,34]. In summary, in scrib−/− benign tumors, the JNK anti-tumorigenic effect consists mainly in the induction of apoptosis while JNK also contributes to differentiation inhibition and degradation of the basement membrane (Figure 2A). It is worth mentioning that JNK signaling does not promote the CyclinE-mediated overproliferation phenotype or cell morphology alterations [69].
In RasV12; scrib−/− tumors, JNK signaling is activated in patches of cells within the tumor and migrating cells that left the primary site [7,9,69]. In these cells, JNK promotes the upregulation of MMP1 and the actin cross-linking protein Cheerio necessary for basement membrane degradation and cytoskeleton remodeling, respectively, together leading to tumor cell migration [7,9,10,70]. It also promotes Paxillin expression, whose function has not been further investigated [69]. Surprisingly, it has been reported that JNK activation in this context is dependent on Kay, but not on its conventional partner Jra [10]. Kay then triggers the expression of the TFs Ftz-f1 and Ets21c (which it associates with), which are essential mediators of the JNK-driven malignancy [54,71,72]. JNK signaling also represses photoneuron specification through the BTB Zinc finger proteins (BTBZF) Chinmo and Abrupt [73]. The JNK target Cheerio also contributes to differentiation inhibition while enhancing proliferation [70]. The switch from anti- to protumorigenic function of JNK lies in the inability of RasV12; scrib−/− cells to undergo apoptosis although caspases are activated [74]. This is due to the fact that RasV12 both downregulates hid [75] and represses Hid activity [42]. At the same time, RasV12 boosts proliferation through the PI3K pathway and Myc [76]. To conclude, in RasV12; scrib−/− cells, JNK promotes invasion (basement membrane and cell motility), differentiation inhibition, and most likely proliferation, but it cannot trigger apoptosis (Figure 2C).
Because JNK signaling is such a strong driver of RasV12; scrib−/− neoplasia, it is crucial to understand how the pathway is activated.

3.1.4. JNK Signaling Upstream Mechanisms in scrib−/− Versus RasV12; scrib−/− Tumors

Whereas Eiger has been reported to be the primary activation signal upstream of JNK in scrib−/− cells [33], several studies report a more complex and diversified upstream network for JNK signaling RasV12; scrib−/− cells.
RasV12; scrib−/− cells have long been thought to activate JNK signaling only through Eiger as eiger loss of function impairs tumor growth. Both Grnd and Wgn seem to be involved [7,55,63,77], although Wgn appears to have a minor contribution to JNK signaling activation compared with Grnd [55,77]. Most probably because of technical issues, the other components leading to JNK activation have only been validated in RasV12; lgl−/− eye disc clones (who display a similar neoplastic phenotype as RasV12; scrib−/− tumors [11]). It consists of Hep, Tak1, and dTraf2 (an adapter protein that activates Tak1) [55]. Several additional players have recently been identified. For instance, the novel JNKK MKK3 (Map Kinase Kinase 3 aka Licorne) positively modulates JNK activation in parallel of Hep and promotes invasion of tumor cells [78]. Additionally, the nonreceptor tyrosine kinase Scr42a activates the dUev1/ben ubiquitin-conjugating enzymes complex upstream of Traf2 and further contributes to JNK signaling [79,80]. Interestingly, overexpression of Src42a in RasV12 tumors promotes neoplastic growth through JNK signaling and PI3K signaling, confirming the link between Src42a and JNK signaling in tumorigenesis [81]. Of note, another non-receptor protein tyrosine kinase called FER has also been identified in the wing disc, where it directly phosphorylates Bsk and promotes migration of scrib knock-down (scribKD) cells [82]. Similarly, the JNKKK Wallenda (Wnd) binds to Rho1 in order to modulate migration in the same system. [63,67] (Figure 2C). Whether FER or the Wnd/Rho1 complex also contribute to neoplasia in the RasV12; scrib−/− eye disc tumors still need to be elucidated. Notably, the nature of the activating signal upstream of MKK3, Src42a, FER, or Wnd has not yet been addressed. Because in other systems, JNK signaling can integrate cues from various processes such as oxidative stress, novel JNK modulators within RasV12; scrib−/− tumor cells might be uncovered in the future [77,83].
Finally, in RasV12 clones, the cells of the edge display basal levels of JNK signaling. Thus, although competent for JNK activation, RasV12 cells seem to refrain from activating JNK in all cells. Autophagy and ROS inhibition are thought to lower JNK signaling in this context [84]. Notably, JNK ectopic activation turns RasV12 benign tumors into invasive tumors [10]. Similarly, in human, JNK, and H-RasV12 also cooperate to promote the invasiveness of human mammary cells. This is of great relevance as Her2+ breast cancer is characterized by both Ras/Erk signaling activation and JNK signaling signature [58].
Taken together, these studies reveal that JNK signaling lies at the heart of the cooperation between RasV12 and the polarity deficiency to drive malignancy in Drosophila and that it seems conserved in human cancer.

3.2. JAK-STAT: A Cytokine-Triggered Signaling Pathway That Amplifies Tumor Growth

The evolutionary conserved JAK-STAT (Janus Kinase and Signal Transducer and Activator of Transcription) has been implicated in processes as diverse as tissue homeostasis control, sex determination, or immune response. In Drosophila, the JAK-STAT pathway is canonically activated by the binding of the cytokines Unpaired (Upd) 1, Upd2, or Upd3 to the dimerized Dome receptor. Upon binding, Dome triggers the activation of the JAK protein Hopscotch (Hop) anchored to its cytoplasmic domain. Hop autophosphorylates itself and phosphorylates Dome to provide free access for the pathway effector STAT92E (Signal Transducer and Activator of Transcription at 92E) to the docking sites on Dome. In this way, inactive STAT92E is recruited and phosphorylated. Subsequently, active STAT92E forms homodimers which translocate to the nucleus. There, STAT92E binds to DNA palindromic sequence. It recruits transcriptional activators and promotes the transcription of many targets, including cyclinD, antimicrobial peptides, or proto-oncogenes such as draf (Reviewed in [85]).
In RasV12; scrib−/− tumors, JAK-STAT signaling is widely activated, and the three Upd ligands are overexpressed (Figure 2C). Inhibition of the JAK-STAT pathway by the overexpression of a dominant negative form of Dome (DomeDN) drastically reduces tumor growth and abrogates invasion. Interestingly, loss of function of stat92e only mildly impairs tumor growth and invasion, suggesting that other uncharacterized JAK downstream effectors might contribute to malignant transformation [86]. Moreover, overexpression of Upd in RasV12 clones is sufficient to induce massive tumor growth and invasion [86], demonstrating that RasV12 can also cooperate with Dome-JAK-STAT to drive malignancy. This observation prompted a few labs to investigate whether the JAK-STAT pathway is linked to JNK signaling in RasV12; scrib−/− tumors.
Indeed, in wing discs homozygous for scrib−/− or bearing dlgKD clones, JNK signaling promotes JAK-STAT activation [86,87]. Moreover, in RasV12, dlgKD eye disc tumors (who display a similar neoplastic phenotype as RasV12; scrib−/− tumors [11]), the JNK downstream effector Fos associates with the TF Ets21c to upregulate upd1 [71]. Thus, although not fully demonstrated, the common view is that JNK signaling activates the JAK-STAT pathway in RasV12; scrib−/− eye disc tumors through autocrine Upd production.
It has also been demonstrated that JAK-STAT activation in the RasV12; scrib−/− tumors increases ROS levels necessary for tumor growth [54,74]. Nevertheless, the identity of the STAT92E targets in this tumor model remains largely unexplored.

3.3. The Salvador–Warts–Hippo Pathway: When the Brake Breaks

The Salvador–Warts–Hippo (SWH) pathway coordinates mechanical inputs with organ homeostasis. It integrates various parameters such as cell polarity, cell contact, and G-protein coupled receptor signaling to fine-tune cell proliferation and survival. Defects of the SWH pathway are associated with poor prognosis in many cancers and are thought to help cancer cells escape cell ‘contact inhibition’ [88,89].
Although the main target of the SWH pathway, YAP (Yes-Associated Protein), was identified first in yeast and mammals, its functions and regulation remained unclear until the discovery of its Drosophila ortholog Yorkie (Yki). Indeed, the upstream regulators of Yki were identified first in Drosophila, contributing greatly to the understanding of the conserved regulation and functions of YAP in mammals both during development and tumorigenesis (reviewed in [89]).
The core of the SWH pathway consists of three interacting proteins: Hippo, Salvador, and Warts. The Hippo (Hpo) Serine Threonine kinase phosphorylates its adaptor protein Salvador (Sav) to form an activated Hpo-Sav complex. This complex phosphorylates and activates the Serine Threonine kinase Warts (Wts) and its adaptor protein Mob As Tumor Suppressor (Mats). Together, they form the activated Wts–Mats complex, which inactivates through phosphorylation the Yes-associated protein ortholog Yki. Yki is a transcriptional co-activator that, when phosphorylated, is kept in the cytoplasm and sent for degradation through its association with 14-3-3 proteins. When its phosphorylation state is low, Yki enters the nucleus. There, it associates with the transcription factor Scalloped (although other Yki partner TFs exist) and activates the transcription of many target genes involved in cell proliferation, like cyclinE, and in cell survival, like diap1 (Death-associated Inhibitor of Apoptosis 1).
Thus, by repressing Yki activity, the SWH pathway is a central negative regulator of proliferation and survival.
Upstream activators of the SWH pathway are key players of cell junction and adhesion. For instance, a subapical complex composed of the FERM domain proteins Merlin and Expanded and the adaptor protein Kibra integrates cell contact and polarity signals. Consequently, this complex binds to the core components of the SWH and promotes sequestration of Yki in the cytoplasm. Similarly, the proto-cadherin Fat modulates the pathway positively.
Scrib−/− clones in the eye disc display an overproliferation phenotype that is not controlled by JNK signaling. Instead, overproliferation results from the inhibition of the SWH pathway by aberrant aPKC activity due to scrib loss of function [26,90] (Figure 2A). Similarly, SWH downregulation is also observed in lgl−/− clones in the eye disc and the wing disc and provokes CyclinE and Diap1 upregulation [26,91]. Unexpectedly, suppression of JNK signaling within scrib−/− eye disc clones enhances Yki elevation, suggesting that JNK restricts tumor growth through SWH signaling [90,92]. Thus, in scrib−/− eye disc clones, Yki activity, promoted by aPKC and repressed by JNK, stimulates tumor cell proliferation. Contrary to the eye disc tumors, in lglKD and dlgKD wing disc tumors, Yki-induced proliferation is promoted by JNK signaling [93]. This significant difference underlines the importance of being cautious while comparing the loss of cell polarity-induced tumors that develop in different tissues.
In RasV12; scrib−/− tumors, high Yki/Sd activity levels promote tumor proliferation and invasion [70,90,94,95] (Figure 2C). Contrary to scrib−/− clones, JNK inhibition, does not affect the expression of Yki targets in RasV12; scrib−/− tumors [73,92], although the JNK target Cher is necessary for the expression of the Yki target expanded [70]. Unexpectedly, targets of Sd within RasV12; scrib−/− tumors are predicted to include the main effectors of the JAK-STAT pathway (stat92E) and JNK pathway (fos, the Jra partner atf3 and ftz-f1) [95] suggesting that Yki might instead act upstream of both JNK and JAK-STAT signaling. Indeed, Yki inhibition in RasV12; scribKD wing disc tumors suppressed JNK signaling and tumor growth [94]. This is puzzling because JNK signaling is upstream of the SWH pathway in scrib−/− eye disc clones. Moreover, several examples of SWH inhibition by the JNK pathway have been reported in both wild type and tumor-bearing wing discs [93,96] and eye discs [97]. Finally, JNK signaling activation in RasV12 eye disc clones increases Yki activity [97,98], reinforcing the idea that JNK signaling is upstream of Yki. Further work would be required to address the exact relationship between the JNK pathway and the SWH pathway. aPKC might be another potential inducer of Yki activity in RasV12; scrib−/− tumors, given its tumor-promoting function in scrib−/− clones. However, aPKC inhibition does not impact RasV12; scrib−/− tumor growth and invasion [69]. Recently, it has been demonstrated that in presence of RasV12, JNK inhibits Wts activity and promotes Yki activity in the eye disc. However, this mechanism has not yet been demonstrated in RasV12; scrib−/− tumors [99].
To conclude, the proliferation, differentiation inhibition, survival, and invasion of RasV12; scrib−/− cells are governed by a cocktail of three main signaling pathways: the JNK, the JAK-STAT, and the SWH pathway (Figure 2). Chromatin profiling and transcriptomic analysis of RasV12; scrib−/− eye disc tumors have validated the importance of these three pathways for reshaping the malignant cells genomic and transcriptomic landscapes. Remarkably, they conclude that less than ten transcription factors account for the neoplastic transformation of RasV12; scrib−/− [95,100]. The JNK transcription factor Fos and the JAK-STAT transcription factor Stat92E are predicted to target 70% and 10% of the tumor-induced regulatory regions, respectively [100]. Nevertheless, the hierarchy and interdependency between them remain unresolved.
Because RasV12; scrib−/− tumor cells express the secreted cytokines Upd1, Upd2, and Upd3, as well as the TNF homolog, Eiger, several studies have focused on how the tumor cells affect their neighboring wild type cells.

4. Microenvironmental Cells Control Tumor Growth in Various Ways

The Drosophila single insult scrib−/− and the RasV12; scrib−/− cooperative oncogenesis models have both been instrumental in understanding how the microenvironment affects cancer progression. Indeed, both anti- and pro-tumorigenic local interactions have been described. Here we will review the basic principles and examples of how the microenvironment can assume both these roles.

4.1. Scrib−/− Model—A Tumor Suppressive Role of the Microenvironment through Cell Competition

Landmark studies in Drosophila have identified a critical “surveillance mechanism” termed cell competition (reviewed in the same Issue [101]). Patches of cells with detrimental mutations, called losers, are actively eliminated by wild type neighbors at the clonal border, termed winner cells [60,102]. This phenomenon ensures that abnormal cell populations do not colonize and dominate the tissue and were later shown to be crucial for suppressing polarity-deficient tumors [34]. For instance, scrib−/− clones are efficiently removed from the epithelium during development, so that very few cells survive to adulthood. Strikingly, when the whole tissue is mutant for scrib or when surrounding cells are genetically ablated, apoptosis in scrib−/− cells is suppressed, and they develop into malignant tumors [34,103]. This observation demonstrates the key tumor-suppressive role of wt neighboring cells.
How do wt cells eliminate their oncogenic insulted neighbors? Indeed, the elimination of these tumors is genetically encoded by a set of molecular programs, which results in their commitment to the loser status, impaired proliferation, induced entosis, and ultimately death.
As described above, the stress pathway JNK is activated in scrib−/− clones, where it triggers apoptosis [33,34]. JNK additionally downregulates Yki activity preventing overproliferation [92]. The dampening of Yki activity is also achieved through the serine protease inhibitor Serpin 5 (Spn5) [104] (Figure 3A(a)). Spn5 inhibits Toll signaling by preventing the activation of the Toll ligand Spaetzle [104]. spn5 knock-down in the microenvironment provokes high levels of toll signaling in scrib−/− clones (Katsukawa et al., 2018). Ectopic activation of Toll signaling turns losers into winners and induces the death of wt microenvironmental cells [104], a phenomenon called supercompetition [104]. These scrib−/−, Toll-high converted winners also have upregulated Yki activity [104], a known trigger of supercompetitor status [105]. Thus, the lowering of Yki activity causes the elimination of polarity deficient clones both by toning down the tumor cell proliferation and preserving the tumor suppressive microenvironment.
An exciting question emerging from these studies is: how do healthy epithelial neighbors sense the presence of transformed cells and execute their elimination?
To restrict scrib−/− tumor growth, the binding of the Receptor tyrosine phosphatase Ptp10D in tumor cells to its surface ligand-protein Sas in wt neighboring cells is critical (Figure 3A(b)). Interestingly, when in contact, the scrib−/− cells and their healthy neighbors perform an apical-lateral translocation of the receptor Ptp10D and the ligand Sas, respectively. Failure in establishing this Ptp10D-Sas crosstalk results in elevated Epithelial Growth Factor Receptor (Egfr) signaling activity, which then cooperates with JNK to drive overgrowth of scrib−/− tumors [106], reminiscent of the RasV12; scrib−/− oncogenesis. In line with this finding, in mammals, the EGFR is negatively modulated through its dephosphorylation by the orthologs of Ptp10D [107,108].
JNK signaling in the microenvironment is also part of the crosstalk between the tumor cells and their neighbors. Indeed, JNK activity is not only upregulated in the scrib−/− tumor but also in the surrounding wt cells [57]. Surprisingly, this JNK activation in the microenvironmental cells is required for tumor suppression by entosis (defined as the engulfment of tumor cells by wt cells) [109]. In these wt neighbors, JNK upregulates the receptor protein-tyrosine kinase Pvr (PDGF- and VEGF-receptor related) and its downstream effectors, the GTPase regulators ELMO (engulfment and cell motility) and Mbc (Myoblast city) [57] (Figure 3A(c)). ELMO and Mbc are thought to form a guanine-nucleotide exchange factor complex which positively regulates Rac1 [110], which is believed to stimulate cytoskeletal dynamics. This should then orchestrate loser cell phagocytosis and ultimately engulfment, and interestingly has been shown to be involved in another cell competition context [111]. Of note, JNK activation in wt border cells does not induce apoptosis [57], contrary to the tumor compartment. A possible explanation for this observation is that Yki activity also seems upregulated in the surrounding neighbors [92]. Then, one of the main targets of Yki, diap1, might promote survival of the wt cells at the border of the tumor despite high JNK activity. Another lead for explaining the survival of border untransformed cells could be the generation of ERK waves. Indeed, it has been shown recently that ectopic induction of apoptosis in epithelia composed of wt cells triggers ERK waves in the surrounding non-dying cells, which supports survival through inhibition of caspases [112]. Therefore, it is plausible that a similar mechanism takes place in the surrounding wt cells of a scrib−/− dying tumor. Future experiments will test whether these phenomena could trigger the conditional survival of the neighboring cells to the JNK stress pathway. To conclude, microenvironmental cells restrict scrib−/− tumor growth by lowering their proliferation through the Ptp10D-Sas axis, eliminating tumor cells by entosis through Pvr signaling, and actively surviving to aberrant JNK signaling through Yki elevation.
Overall, these studies showcase the critical importance of the local host cells, where several genetic programs are put in place to halt tumor progression.

4.2. RasV12; scrib−/− Model—A Microenvironmental Pro-Tumorigenic Role Switch in a Cooperative Scenario

4.2.1. Undead RasV12; scrib−/− Tumor Cells Exploit Apoptosis-Induced Proliferation, ROS, and Inflammation

Upon the elimination of unfit cells from the tissue, surrounding surviving cells overproliferate to preserve tissue homeostasis. This safeguarding mechanism is known as apoptosis-induced proliferation (AiP). The apoptosis of a cell within a Drosophila imaginal epithelium leads to the generation of reactive oxygen species (ROS) and mitogens, which stimulate the proliferation of the neighboring cells [113]. Indeed, AiP is thought to be responsible for the normal appearance of the adult eyes in which scrib−/− tumors formed and were progressively eliminated (Wu et al., 2010b). In detail, the downregulation of Upd and Stat92E in the scrib−/− tumor and its microenvironment, respectively, account for a visible reduction of the adult eye. This observation shows that a JAK-STAT communication axis between both compartments is essential to execute AiP and replenish tissue cell loss [86].
Nevertheless, the large majority of our knowledge underlying the molecular mechanisms behind AiP comes from the study of “undead” cells. In this model, cells are fated genetically for cell death through the upregulation of the pro-apoptotic factors Reaper and Hid. Typically, they trigger the activation of the initiator caspase Dronc through Diap1 inhibition. Then, Dronc activates the effector Caspases Dcp-1 (Decapping protein 1) and DriCE (Death-related ICE-like caspase), leading to apoptosis. Upon coexpression of the effector caspase inhibitor, p35, apoptosis cannot proceed, leaving the cell in an undead state [113].
In “undead” cells, Dronc can activate the production of extracellular ROS through Duox, which alarms hemocytes to mount an inflammatory response through Eiger. This is suggested to promote JNK activation in epithelial cells, mitogen expression, and AiP [114].
Interestingly, RasV12; scrib−/− tumors behave much like the classical “undead” cells and exploit a similar mechanism [74]. Strikingly, RasV12; scrib−/− tumors strongly accumulate ROS, most probably because of impaired mitochondrial function [54] and Caspase activity [74] as well as they attract hemocytes (Figure 3B(a)). Upon the downregulation of both extra and intracellular ROS production, JNK signaling is inhibited, tumor growth is largely impaired, and adult survival is restored. These findings demonstrate how critical ROS are for malignant transformation [74]. Overall, this study suggests tumors hijack the AiP mechanisms for ROS generation through Caspases to elevate JNK levels [74], which is essential to drive malignancy in this tumor model.

4.2.2. Non-Autonomous Autophagy Fuels Tumor Growth

Due to their aggressive proliferative growth and migratory behavior, neoplastic tumors have a high anabolic demand. Neighboring healthy cells could constitute a valuable source of nutrients. RasV12; scrib−/− tumors induce autophagy in the microenvironment [54] (Figure 3B(b)). Autophagy is a process by which cells can degrade macromolecular components in the lysosome to recycle building blocks in nucleotides, sugars, amino acids, and fatty acids for anabolic or energetic demands. When autophagy is inhibited in the microenvironment, it strongly reduces tumor growth and malignancy in vivo [54,115]. Downregulation of an amino acid transporter Slimfast (Slif) limits the growth of a related cooperative model, RasV12; dlg−/− [54], suggesting that these tumors are highly dependent on nutrients from their environment. In line with this finding, pancreatic tumors are provided amino acids via autophagy from their neighboring stellate cells [116]. In RasV12; scrib−/− eye disc tumors, non-autonomous autophagy (NAA) was shown to depend on the JNK signaling in the tumor and the autocrine activation of JAK-STAT through the Upd cytokines and Dome receptor [54]. However, the exact trigger and sensor for NAA in the microenvironment and whether nutrient mobilization is the sole role of autophagy in this model remains to be discovered.
Overall, although autophagy has been classically regarded as a tumor-suppressive mechanism, this study reinforced the idea that it can play an ambivalent role in cancer [117]. The capacity to be pro-tumorigenic was dependent on an oncogenic cooperation network, highlighting the importance of the tumor mutational context when designing therapeutic approaches to target this pathway.

4.2.3. Tumor Hotspots

A central hypothesis in the field of cancer research is the existence of pro-tumorigenic niches: specific areas in tissues or organs that are more favorable for initial tumor growth or metastasis. Understanding the exact molecular nature of why some zones are more nurturing is of high interest since this could allow the prediction of oncogenic target sites and early cancer detection.
Interestingly, a study in the wing imaginal disc identified a tissue region where large dysplastic lglKD and scribKD tumors (defined as abnormally growing lesions) were consistently generated, as opposed to another area where these overgrowths were not observed. These zones where thus designated tumor “hotspots” and “coldspots”, and they were found in the dorsal hinge and the disc pouch, respectively [118] (Figure 3B(c)).
These areas have two critical differences. Inflammatory pro-proliferative JAK-STAT signaling is active during development in the dorsal hinge. Furthermore, upon the generation of lglKD and scribKD clones, in both regions, extrusion of apoptotic cells to the basal side of the disc was observed. However, only in the hotspot zone, polarity deficient clones extruded apically and generated dysplastic tumors [118].
Indeed, Stat92E was found to be required in the apically extruded clones in the hotspot zone to generate dysplastic large clones. This exciting result shows that endogenous, “pre-patterned” tissue signaling programs can cooperate with tumor cells and drive their growth. Furthermore, through the knock-down of the cytoskeletal regulator rho guanine nucleotide exchange factor 2 (rho-gef2) it is possible to induce both basal and apical extrusion of lglKD and scribKD clones in the pouch region. However, this is not sufficient to induce large dysplastic tumors in the pouch. This is only possible upon ectopic induction of JAK-STAT signaling combined with induced apical extrusion of clones in the coldspot zone [118].
To conclude, the endogenous tissue signaling and the polarity of cell extrusion can vary spatially, providing certain areas with increased risk for tumorigenesis. This study demonstrated an important principle, where cooperative scenarios promoting tumor growth can be generated through the interaction of gene mutations and additionally endogenous wildtype tissue programs installed during development and potentially homeostasis.

4.2.4. Interclonal Cooperation

Classically, the microenvironment is classified as all non-transformed cells surrounding the tumor and the extracellular matrix. Additionally, in the vicinity of a tumor cell, other cancer cells with similar or different mutant backgrounds can exist and potentially establish a communication axis to influence each other’s fate. Drosophila offers a platform for the dissection of the molecular underpinnings behind such clonal interactions and the opportunity to identify new basic molecular principles.
A previous study examined the behavior of separate RasV12 and scrib−/− clones generated individually in the same eye antennal disc (RasV12//scrib−/−) [86]. Strikingly, these clones cooperate, resulting in strong overgrowth and invasion into the ventral nerve cord (Figure 3B(d)). These tumors are composed mainly of RasV12 cells. Like in double mutant RasV12; scrib−/− tumors, a reporter for JAK-STAT showed high activity in tumors, and the upd ligands were highly expressed. Inhibiting JAK-STAT signaling through DomeDN expression in RasV12 cells facing scrib−/− cells completely inhibited tumor growth and invasion, demonstrating the critical role of inflammatory pro-proliferative signals in this interclonal cooperation scenario. Furthermore, inhibiting JNK in RasV12 cells, but not in neighboring scrib−/− clones results in a substantial reduction in growth and inhibition of malignancy. Moreover, concomitant expression of Upd and BskDN (a dominant negative form of Bsk) in RasV12 clones recapitulated the phenotype observed in RasV12//scrib−/−. RasV12 clones have low JNK levels, and thus these results suggest that JNK signaling can propagate from scrib−/− to RasV12 tumors within the same disc. This then drives the upregulation of JAK-STAT and sequentially the overgrowth and malignancy of RasV12 clones [86].
To prove that JNK signaling can propagate, the authors carried out an elegant experiment in the wing imaginal disc. Here, the inhibitor of JNK, Puckered, was overexpressed medially in the patched domain. Wounding, a known trigger of JNK, was then performed in only one side of the disc. Interestingly, as compared to control, JNK activity did not propagate beyond the patched stripe overexpressing Puckered [86]. This demonstrated that JNK indeed propagates and that this propagation requires JNK activity. However, others have shown that JNK propagation to the host cells from dlg−/− tumors are independent of JNK [57]. Possibly, in a tumorigenic vs. wounding scenario, additional mechanisms are put in place to trigger a host response, but this signal still remains undiscovered.
Thus, non-autonomous tumor interactions are not restricted to the wt neighbors but can additionally be established with other tumors in their vicinity. Furthermore, Wu et al., underlined that oncogenic cooperation does not necessarily need to be generated within the same cell population but can be established from individually adjacent compartments to achieve a malignant outcome.

5. Long-Range Interactions between the Tumor and Its Host Are Critical for Tumor Growth and Invasion

Tumor cells not only communicate with their direct neighbors but also initiate cross-talk with distant organs. This section reviews how the RasV12; scrib−/− tumors lessen ecdysone production in the prothoracic gland to halt larval to pupal transition-pupation. Then we focus on the enigmatic contribution of the inflammation response mediated by tumor-associated immune cells and NFκB signaling in the tumor. Finally, we discuss the contribution of tumor-induced organ wasting.

5.1. The Tumor Halts the Larval Development and Gains More Time to Grow

RasV12; scrib−/− tumor-bearing larvae linger, meaning that instead of pupariating at day 6 and undergoing metamorphosis into an adult at day 10, they extend their larval life up to 15 days. During this aberrant larval period, tumor-bearing larvae keep growing (along with their tumor), progressively lose motility and ultimately die [11].
The developmental arrest of RasV12; scrib−/− tumor-bearing larvae was the first evidence that these tumors have systemic effects and impair the developmental progression and physiology of the whole organism.
Drosophila develops through three successive larval stages (namely L1, L2, and L3) before initiating pupation, during which metamorphosis reshapes the larval structures into finalized adult organs. Molts, pupation, and metamorphosis are tightly controlled by regulated bursts of production of the steroid hormone Ecdysone by the prothoracic gland [119]. Ecdysone production is finely balanced by systemic growth inputs which ensure proper organism size and symmetry control. However, upon damage of larval structures, Ecdysone production (and consequently development) is put on hold while systemic growth is slowed down until the wound is repaired. The insulin-like peptide Dilp8 is the main messenger that triggers developmental arrest upon tissue damage. For instance, damaged imaginal discs overexpress and secrete Dilp8 in the hemolymph [120,121]. Dilp8 binds to the relaxin receptor Lgr3 (Leucine-rich repeat-containing G-protein-coupled receptor 3) in the GCL neurons in the brain. These neurons activate a circuit that ultimately abrogates Ecdysone production in the prothoracic gland [122,123,124,125].
An old concept in cancer research is that tumors can be considered as wounds that do not heal. In line with this idea, RasV12; scrib−/− tumor-bearing eye discs (and other imaginal disc tumor models) overexpress and secrete Dilp8 [72,121,126]. As a result, Ecdysone production decreases, and pupation is blocked. How Dilp8 expression is induced by RasV12; scrib−/− eye discs tumors is not entirely clear. Interestingly, the JNK and SWH pathways are the primary potential regulators of Dilp8 expression. Indeed, Dilp8 overexpression in RasV12; scrib−/− eye disc tumors is strongly down-regulated upon loss of function of the main JNK effectors fos, ets21c, and ftz-f1 and is accompanied by a rescue of both the pupation defect and adult survival [72].
Additionally, it has recently been shown that elevated RasV12 in the eye discs leads to increased JNK signaling and further dilp8 upregulation in late larval stages. Although not able to prevent pupation alone, RasV12-mediated Dilp8 elevation lowers Ecdysone production so that after pupation, metamorphosis cannot proceed and the pupae die [49]. It is thus tempting to speculate that in RasV12; scrib−/− eye disc tumors JNK signaling would be a major regulator of dilp8 expression and Ecdysone signaling impairment. However, a study in the wing disc strongly supports the idea that although responsive to JNK signaling, dilp8 transcriptional activity is mainly controlled by the Yki partner Sd, which binds to specific sites in its enhancer region [127]. Given the crucial role of Yki for RasV12; scrib−/− tumor proliferation and invasion, future work might also uncover its contribution to the pupation defects together with JNK.
Nevertheless, JNK signaling inhibition in RasV12; scrib−/− tumors restores pupation and improves adult survival, demonstrating the involvement of this pathway in the developmental delay phenotype [72]. Noteworthy, this phenomenon might also occur in a Dilp8-independent manner. Like, Dilp8, Upd3 has recently been implicated in inducing a developmental delay in a chromosomal instability (CIN)—driven wing disc tumor model [128]. In this model, secreted Upd3 acts as a systemic signal that promotes a developmental delay through JAK-STAT signaling and bantam microRNA upregulation in the prothoracic gland. Consequently, Ecdysone biosynthesis decreases, and the metamorphosis onset is delayed [128]. Because upd3 is strongly upregulated in RasV12; scrib−/− tumors in a JNK-dependent manner, a similar Ecdysone-inhibiting mechanism might occur due to Upd3 expression and contribute to the lingering phenotype [86].

5.2. Innate Immune Response: The Tumor as a Wound That Never Heals

RasV12; scrib−/− not only triggers damage-like development arrest, it also instigates a strong systemic inflammation response.

5.2.1. The Enigmatic Role of the Tumor-Associated Hemocytes

In Drosophila, the immune response is partially mediated by the macrophage-like cells called hemocytes. They are a primary line of defense against fungi, bacteria, and yeast. The three types of hemocytes: the plasmatocytes, the lamellocytes, and the crystal cells, execute phagocytosis/coagulation, encapsulation, and melanization, respectively [129]. Upon infection, hemocytes activate the fat body’s systemic immune response through the cytokine Upd3. The fat body (Figure 1A) (whose function is similar to the mammalian liver) activates the Toll or IMD NFkB pathways and massively secrete pathogen-tailored anti-microbial peptides into the hemolymph [129].
In RasV12; scrib−/− tumor-bearing larvae (as well as in scrib−/− and dlg−/− larvae which display neoplastic growth in the imaginal discs), the number of circulating hemocytes increases due to higher hemocyte proliferation. Plasmatocytes adhere to the surface of RasV12; scrib−/− tumors where the basement membrane is degraded, similarly to injured disc, emphasizing the notion that these tumors are perceived as wounds by the organism [7,8,70,74,77,130]. Investigating the role of these tumor-associated hemocytes (TAH) is technically and genetically challenging. Several studies involving demanding techniques such as tumor transplantation or hemocytes transfusion have elegantly addressed this question. Unfortunately, they do not reach the same conclusions and report both a pro and an anti-tumorigenic role of the RasV12; scrib−/− TAHs.
To follow tumor growth over a long time inside a host (instead of 15 days in the larva), a common technique consists of cutting and transplanting tumor pieces into the abdomen of an adult female fly [131]. When transplanted, RasV12; scribKD wing disc tumors attract adult hemocytes which restrict tumor growth [77]. Similarly, neoplastic tumors from whole scrib−/− eye discs as well as dlg−/− eye discs (which display neoplastic growth) grow better when the TAHs are killed in the larva [8,130]. Interestingly, the TAHs associated with eye disc tumors in dlg−/− mutant larvae, secrete both Egr and Spz. In this context, Egr triggers local tumor cell apoptosis suggesting an anti-tumorigenic role of TAHs in scrib−/− and dlg−/− mutant larvae [8,130,132]. By contrast, in RasV12; scrib−/− eye disc tumor-bearing larva, the TAHs also secrete Eiger and contribute to JNK signaling through MMP1 expression within the tumor arguing for a tumor-invasion supportive role [7] although no observation on tumor growth has been made. JNK (Cher) and JAK-STAT (STAT-GFP) signaling have been observed in TAHs but not further investigated [8,70]. These opposite potential roles of TAHs might be inherent to the type of tumors they associate with. Indeed, JNK signaling is protumorigenic only when RasV12 is expressed and inhibits apoptosis. It is, therefore, tempting to speculate that in RasV12; scrib−/− tumors, TAHs might be tumorigenic through the promotion of JNK signaling in the tumor, whereas TAHs are antitumorigenic in scrib−/− and dlg−/− larvae for the very same reason. Future investigations would be required to elucidate these questions.
A related question is how TAHs are recruited to the tumor. RasV12; scrib−/− tumors produce ROS in a JNK-dependent manner. ROS are particularly enriched where the basement membrane is degraded through the activity of the transmembrane NADPH oxidase Duox [74]. In RasV12; scrib−/− tumors, ROS production inhibition both intracellularly and extracellularly abrogates TAHs recruitment. Interestingly, a similar hemocyte recruitment mechanism is observed in the eye discs bearing undead cell clones suggesting that the RasV12, scrib−/− tumor growth relies partly on AiP mechanisms and that hemocytes are part of this process [133].

5.2.2. NFκB Activation in the Tumor Triggers a Humoral Inflammation Response

Another part of the inflammation response is characterized by the NFκB-mediated elevation of AMPs in the hemolymph. TAHs in dlg−/− mutant larvae triggers a long range Toll-mediated inflammation response in the fat body by secreting Spz [130]. Indeed, Spz activates Toll signaling in the fat body and trachea which results in the secretion of the anti-microbial peptide (AMP) Defensin. Interestingly, secreted Defensin attach to phospatidylserines that are exposed at the surface of the tumor in an Eiger-dependent manner ultimately provoking tumor cell death [132]. Moreover, a recent study has shown that RasV12; scrib−/− tumor-bearing larvae have elevated levels of the AMP Drosomycin, also a target of the Toll pathway, confirming that the presence of the tumor within the larva induces a systemic inflammation response [134]. Interestingly, bacterial but not fungal infection of RasV12; scrib−/− tumor-bearing larvae can modulate this inflammation and results in weaker tumor growth, suggesting that inflammation restricts tumor growth [134]. Recently, an elegant RNAi-mediated knock-down screen has been performed in vitro on RasV12; scrib−/− cells challenged with a scratch assay to characterize novel invasion regulators. Toll6, one of the receptors of the Toll pathway, is enriched in the tumor and governs the trajectory of invasive RasV12; scrib−/− cells toward specific organs (aka organotropism), through its binding to the ligand Spz. Strikingly, only the Spz-expressing larval organs are susceptible to secondary tumor formation [53] suggesting that NFκB signaling has an unexpected role in cancer invasion. Finally, a recent study reported that the negative NFκB regulator Cactus is also enriched in RasV12; scribKD wing disc tumors. In this context, Cactus is proposed to contribute to tumor growth independently of the NFκBs but rather through the promotion of Yki signaling. Yet, additional studies would be required to validate and assess the relevance of this unconventional mechanism [94].
To conclude, a cohort of studies report that RasV12; scrib−/− tumors initiate an inflammation response within the larvae with an elevation of specific AMPs and the potential involvement of components of the NFκB pathways. This is thought to modulate tumor growth and invasion. Additionally, hemocytes are attracted to the tumor, and the nature of this crosstalk involves ROS production and Eiger secretion [7,74]. However, the function of this association still needs to be clarified in the different tumor models.
Nevertheless, these observations emphasize that the RasV12; scrib−/− tumor model is also suited for studying the role of innate immunity in cancer. Given the high degree of conservation of the NFκB pathway, future findings might contribute to understanding the role of innate immunity in human tumorigenesis.

5.3. Wasting through Systemic Induced Autophagy Is Needed for Tumor Growth

RasV12; scrib−/− tumors can grow in adult fly hosts upon allograft transplantation and eventually kill the host. RasV12; scrib−/− tumors lead to abdominal bloating, shrinkage of the fat body, altered mitochondrial morphology of muscles, and ovary wasting [11,126]. These changes are reminiscent of systemic effects during cancer cachexia in humans.
The functional relevance of this phenotype is still unclear; however, it seems to be a shared characteristic among several neoplastic tumors [135,136].
RasV12; scrib−/− tumor-bearing larvae suffer weight loss, impaired motility and decreased feeding over time [115,137]. The energy-storing organs, such as the muscle and the fat body in the larva, progressively deteriorate over time so that many muscles shrink or rupture, and adhesion of fat body cells decreases and attains a “pearl-like” phenotype during the late stages of tumorigenesis [115,126,137]. The reduction of energy storage accompanies structural defects in these organs. For instance, glycogen storage is significantly reduced in the muscle and the fat body and accompanied by mitochondrial defects. In contrast, lipid storage as triacylglycerides in the fat body seems to be stable, as lipid-free cytoplasm shrinks and lipid droplet aggregation ensues [115,126]. These changes correlate with an increase of amino acid and sugar levels in the larval hemolymph (insect blood), a phenotype shared with a high-sugar diet-induced RasV12, csk−/− larval tumor model [115,135,136,137,138]. The hyperglycemia and amino acid elevation in the hemolymph appear before the onset of alterations in feeding. This observation suggests that nutrient mobilization from wasted organs in the RasV12; scrib−/− tumor-bearing larvae is mainly attributed to a long-range action of the tumor rather than starvation [115,137]. Collectively, this cohort of alterations of the organ structure and the systemic metabolism defines a wasting phenotype that is also observed in cancer patients who suffer from cachexia, establishing Drosophila as a relevant model to investigate the cachexia-associated mechanism (reviewed in [139]).
This wasting phenotype raises two main questions. (1) How does the tumor induce organ wasting? (2) Does organ wasting contribute to tumor growth?

5.3.1. Tumors Secrete Wasting-Inducing Ligands

Initial studies aimed to answer the first question. Transplanted RasV12; scrib−/− tumors secrete the insulin growth factor binding protein (IGFBP) homolog ImpL2 which mediates organ wasting [126]. The mechanism through which tumor-secreted ImpL2 mediates ovary wasting was further detailed using an additional Yki-induced adult gut model [136]. In both models, ImpL2 executes muscle wasting by decreasing systemic Insulin Receptor-PI3K-AKT signaling, and downregulating energy production genes involved in glycolysis, pyruvate metabolism and oxidative phosphorylation in the muscle. Interestingly, the ImpL2 regulatory region possesses binding sites for both Fos and Scalloped suggesting that in RasV12; scrib−/− eye disc tumors both the JNK and the SWH pathways might control its expression [126]. However, ImpL2 is not the sole mediator of wasting in Drosophila. In the Yki-induced gut tumor model, tumors also secrete the Pvf1 (PDGF- and VEGF-related factor 1) which is thought to trigger adipose and muscle wasting through Pvr/ERK signaling [135]. In the RasV12, csk−/− high-sugar diet-induced tumor model, systemic muscle wasting is neither governed by ImpL2 nor by Pvf1. Instead, the fibroblast growth factor Branchless (Bnl) is secreted by the tumor and mediates both muscle wasting and tumor growth [138]. Thus, it would be interesting to decipher whether Pvf1 or Bnl also contribute to the wasting phenotype induced by RasV12; scrib−/− tumors.
In conclusion, it appears that depending on the tumor type, various systemic signals are generated by the transformed cells in order to orchestrate a common host metabolic reprogramming which results in the progressive wasting of energy storage organs (Figure 4). The exact nature of the crosstalk between the tumor and the targeted organs is still an open question.

5.3.2. Organ Wasting Feeds the Tumor

A long-standing question has been whether tumor-associated organ wasting is an unintended side effect, or whether it serves a specific purpose for tumor growth, be it through cell signaling or nutrient supply. A recently described stable carbon isotope tracing technique has allowed to distinguish whether tumor biomass is derived from food or host tissues. Using this technique, it was demonstrated that RasV12; scrib−/− larval tumors first grow mainly on nutrients from food and progressively incorporates more nutrients from host tissues [115,140]. Autophagy is aberrantly activated in muscles and fat body of RasV12; scrib−/− tumor-bearing larvae suggesting that wasted organs might provide nutrients to the tumor through “organ self-digestion” [54]. Indeed, systemic autophagy inhibition demonstrated that autophagy is required for organ wasting and elevated amino acid and sugar levels in hemolymph that supports tumor growth [115]. Thus, RasV12; scrib−/− tumors consume nutrients originating from wasting organs, although the exact source of nutrients and how tumors activate systemic autophagy and organ wasting is still unclear.

6. Conclusions

The cooperation between the oncogene RasV12 and the loss of the tumor suppressor scribble gives rise to invasive malignant tumors in the Drosophila eye-antennal disc. In this model, tumor cells adopt a disorganized and altered morphology accompanied with the aberrant activation of key signaling pathways (JNK, Yki, Raf/Mek/MAPK, PI3K/AKT and JAK-STAT). This cocktail of mitogenic signals fuels the neoplastic tumor growth which in turn severely alters the tissue integrity through dismantling cell adhesion, degrading the basement membrane, counteracting differentiation as well as stimulating cell growth and ectopic proliferation. Finally, the tumor alters the whole-body physiology of the host by impairing hormonal signaling, halting host development, invading and damaging local and distant organs, triggering a systemic inflammation response and consuming the nutrients from the degradation of host tissues. Together, this results in loss of body motility, decrease in feeding and ultimately death of the host.
Unexpectedly, this originally simple in vivo Drosophila tumor model, which has been elaborated upon for nearly two decades, has brought insights into tumorigenesis that involves conserved components, associated with malignant cancer progression and bad prognosis in human cancer (loss of cell polarity, aberrant activation of conserved signaling pathways, ROS production, NFκB inflammatory signaling and organ wasting). Excitingly, the main pathways involved in RasV12; scrib−/− tumorigenesis also turn out to be a common feature of most of the other Drosophila cooperative oncogenic models [141], suggesting that these conserved signaling pathways might also be a common conserved underlying transformation program critical for driving genetic cooperation in human cancers.
Still, many unexplored questions need to be addressed, in particular the nature and functions of the distinct communication signals between the tumor cells and their immediate wild type neighboring cells or distant wasting organs. How does tumors outcompete and possibly consume neighboring cells by supercompetition? What are the mechanisms of metabolic reprogramming of tumor and host? How do tumor cells inflict harm on the host and lead to death? Effectively addressing these and other questions of tumor–host dialogue will require combinatorial genetic manipulation techniques that will allow both the generation of Rasv12; scrib−/− tumors (or other tumor models) and the genetic manipulation of any cellular process in selected organs in vivo.
We predict that the variety of synergistic tumor models combined with novel genetic approaches available in Drosophila, coupled with increasingly sensitive omic approaches will keep on contributing to the acquisition of precious novel and transferrable knowledge and concepts in cancer biology.

Author Contributions

C.D., J.G.T.R. and T.E.R. conceptualized the review. C.D. and J.G.T.R. wrote the first draft and made figures. C.D., J.G.T.R. and T.E.R. edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

C.D., J.T.R. and T.E.R. were supported by the NORWEGIAN RESEARCH COUNCIL (to T.E.R.), grant numbers #262652 and #276070.

Acknowledgments

Figure 3 has been created using BioRender.com accessed on 7 July 2021.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kortlever, R.M.; Sodir, N.M.; Wilson, C.H.; Burkhart, D.L.; Pellegrinet, L.; Brown Swigart, L.; Littlewood, T.D.; Evan, G.I. Myc Cooperates with Ras by Programming Inflammation and Immune Suppression. Cell 2017, 171, 1301–1315.e1314. [Google Scholar] [CrossRef]
  2. Kim, M.P.; Li, X.; Deng, J.; Zhang, Y.; Dai, B.; Allton, K.L.; Hughes, T.G.; Siangco, C.; Augustine, J.J.; Kang, Y.; et al. Oncogenic KRAS recruits an expansive transcriptional network through mutant p53 to drive pancreatic cancer metastasis. Cancer Discov. 2021. [Google Scholar] [CrossRef]
  3. Mirzoyan, Z.; Sollazzo, M.; Allocca, M.; Valenza, A.M.; Grifoni, D.; Bellosta, P. Drosophila melanogaster: A Model Organism to Study Cancer. Front. Genet. 2019, 10, 51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Hanahan, D.; Weinberg, R.A. The hallmarks of cancer. Cell 2000, 100, 57–70. [Google Scholar] [CrossRef] [Green Version]
  5. Bangi, E.; Smibert, P.; Uzilov, A.V.; Teague, A.G.; Gopinath, S.; Antipin, Y.; Chen, R.; Hecht, C.; Gruszczynski, N.; Yon, W.J.; et al. A Drosophila platform identifies a novel, personalized therapy for a patient with adenoid cystic carcinoma. Iscience 2021, 24, 102212. [Google Scholar] [CrossRef] [PubMed]
  6. Richardson, H.E.; Portela, M. Modelling Cooperative Tumorigenesis in Drosophila. Biomed. Res. Int. 2018, 2018, 4258387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Cordero, J.B.; Macagno, J.P.; Stefanatos, R.K.; Strathdee, K.E.; Cagan, R.L.; Vidal, M. Oncogenic Ras diverts a host TNF tumor suppressor activity into tumor promoter. Dev. Cell 2010, 18, 999–1011. [Google Scholar] [CrossRef] [Green Version]
  8. Pastor-Pareja, J.C.; Wu, M.; Xu, T. An innate immune response of blood cells to tumors and tissue damage in Drosophila. Dis. Model. Mech. 2008, 1, 144–154; discussion 153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Srivastava, A.; Pastor-Pareja, J.C.; Igaki, T.; Pagliarini, R.; Xu, T. Basement membrane remodeling is essential for Drosophila disc eversion and tumor invasion. Proc. Natl. Acad. Sci. USA 2007, 104, 2721–2726. [Google Scholar] [CrossRef] [Green Version]
  10. Uhlirova, M.; Bohmann, D. JNK- and Fos-regulated Mmp1 expression cooperates with Ras to induce invasive tumors in Drosophila. EMBO J. 2006, 25, 5294–5304. [Google Scholar] [CrossRef]
  11. Pagliarini, R.A.; Xu, T. A genetic screen in Drosophila for metastatic behavior. Science 2003, 302, 1227–1231. [Google Scholar] [CrossRef]
  12. Halaoui, R.; McCaffrey, L. Rewiring cell polarity signaling in cancer. Oncogene 2015, 34, 939–950. [Google Scholar] [CrossRef]
  13. Siziopikou, K.P. Ductal carcinoma in situ of the breast: Current concepts and future directions. Arch. Pathol. Lab. Med. 2013, 137, 462–466. [Google Scholar] [CrossRef]
  14. Bostwick, D.G.; Cheng, L. Precursors of prostate cancer. Histopathology 2012, 60, 4–27. [Google Scholar] [CrossRef] [PubMed]
  15. Catterall, R.; Lelarge, V.; McCaffrey, L. Genetic alterations of epithelial polarity genes are associated with loss of polarity in invasive breast cancer. Int. J. Cancer 2020, 146, 1578–1591. [Google Scholar] [CrossRef] [PubMed]
  16. Jung, H.Y.; Fattet, L.; Tsai, J.H.; Kajimoto, T.; Chang, Q.; Newton, A.C.; Yang, J. Apical-basal polarity inhibits epithelial-mesenchymal transition and tumour metastasis by PAR-complex-mediated SNAI1 degradation. Nat. Cell Biol. 2019, 21, 359–371. [Google Scholar] [CrossRef] [PubMed]
  17. Williams, E.; Villar-Prados, A.; Bowser, J.; Broaddus, R.; Gladden, A.B. Loss of polarity alters proliferation and differentiation in low-grade endometrial cancers by disrupting Notch signaling. PLoS ONE 2017, 12, e0189081. [Google Scholar] [CrossRef]
  18. Halaoui, R.; Rejon, C.; Chatterjee, S.J.; Szymborski, J.; Meterissian, S.; Muller, W.J.; Omeroglu, A.; McCaffrey, L. Progressive polarity loss and luminal collapse disrupt tissue organization in carcinoma. Genes Dev. 2017, 31, 1573–1587. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Lin, W.H.; Asmann, Y.W.; Anastasiadis, P.Z. Expression of polarity genes in human cancer. Cancer Inform. 2015, 14, 15–28. [Google Scholar] [CrossRef] [Green Version]
  20. Pearson, H.B.; Perez-Mancera, P.A.; Dow, L.E.; Ryan, A.; Tennstedt, P.; Bogani, D.; Elsum, I.; Greenfield, A.; Tuveson, D.A.; Simon, R.; et al. SCRIB expression is deregulated in human prostate cancer, and its deficiency in mice promotes prostate neoplasia. J. Clin. Investig. 2011, 121, 4257–4267. [Google Scholar] [CrossRef] [Green Version]
  21. Zhan, L.; Rosenberg, A.; Bergami, K.C.; Yu, M.; Xuan, Z.; Jaffe, A.B.; Allred, C.; Muthuswamy, S.K. Deregulation of scribble promotes mammary tumorigenesis and reveals a role for cell polarity in carcinoma. Cell 2008, 135, 865–878. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Saito, Y.; Desai, R.R.; Muthuswamy, S.K. Reinterpreting polarity and cancer: The changing landscape from tumor suppression to tumor promotion. Biochim Biophys Acta Rev. Cancer 2018, 1869, 103–116. [Google Scholar] [CrossRef] [PubMed]
  23. Humbert, P.O.; Grzeschik, N.A.; Brumby, A.M.; Galea, R.; Elsum, I.; Richardson, H.E. Control of tumourigenesis by the Scribble/Dlg/Lgl polarity module. Oncogene 2008, 27, 6888–6907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Bonello, T.T.; Choi, W.; Peifer, M. Scribble and Discs-large direct initial assembly and positioning of adherens junctions during the establishment of apical-basal polarity. Development 2019, 146, dev180976. [Google Scholar] [CrossRef] [PubMed]
  25. Choi, J.; Troyanovsky, R.B.; Indra, I.; Mitchell, B.J.; Troyanovsky, S.M. Scribble, Erbin, and Lano redundantly regulate epithelial polarity and apical adhesion complex. J. Cell Biol. 2019, 218, 2277–2293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Grzeschik, N.A.; Parsons, L.M.; Allott, M.L.; Harvey, K.F.; Richardson, H.E. Lgl, aPKC, and Crumbs regulate the Salvador/Warts/Hippo pathway through two distinct mechanisms. Curr. Biol. 2010, 20, 573–581. [Google Scholar] [CrossRef] [Green Version]
  27. Qin, Y.; Capaldo, C.; Gumbiner, B.M.; Macara, I.G. The mammalian Scribble polarity protein regulates epithelial cell adhesion and migration through E-cadherin. J. Cell Biol. 2005, 171, 1061–1071. [Google Scholar] [CrossRef]
  28. Grifoni, D.; Garoia, F.; Schimanski, C.C.; Schmitz, G.; Laurenti, E.; Galle, P.R.; Pession, A.; Cavicchi, S.; Strand, D. The human protein Hugl-1 substitutes for Drosophila lethal giant larvae tumour suppressor function in vivo. Oncogene 2004, 23, 8688–8694. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Dow, L.E.; Brumby, A.M.; Muratore, R.; Coombe, M.L.; Sedelies, K.A.; Trapani, J.A.; Russell, S.M.; Richardson, H.E.; Humbert, P.O. hScrib is a functional homologue of the Drosophila tumour suppressor Scribble. Oncogene 2003, 22, 9225–9230. [Google Scholar] [CrossRef] [Green Version]
  30. Bonello, T.T.; Peifer, M. Scribble: A master scaffold in polarity, adhesion, synaptogenesis, and proliferation. J. Cell Biol. 2019, 218, 742–756. [Google Scholar] [CrossRef] [PubMed]
  31. Thomas, M.; Massimi, P.; Navarro, C.; Borg, J.P.; Banks, L. The hScrib/Dlg apico-basal control complex is differentially targeted by HPV-16 and HPV-18 E6 proteins. Oncogene 2005, 24, 6222–6230. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Germani, F.; Bergantinos, C.; Johnston, L.A. Mosaic Analysis in Drosophila. Genetics 2018, 208, 473–490. [Google Scholar] [CrossRef] [Green Version]
  33. Igaki, T.; Pastor-Pareja, J.C.; Aonuma, H.; Miura, M.; Xu, T. Intrinsic tumor suppression and epithelial maintenance by endocytic activation of Eiger/TNF signaling in Drosophila. Dev. Cell 2009, 16, 458–465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Brumby, A.M.; Richardson, H.E. scribble mutants cooperate with oncogenic Ras or Notch to cause neoplastic overgrowth in Drosophila. EMBO J. 2003, 22, 5769–5779. [Google Scholar] [CrossRef] [PubMed]
  35. Gimple, R.C.; Wang, X. RAS: Striking at the Core of the Oncogenic Circuitry. Front. Oncol. 2019, 9, 965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Ottaiano, A.; Normanno, N.; Facchini, S.; Cassata, A.; Nappi, A.; Romano, C.; Silvestro, L.; De Stefano, A.; Rachiglio, A.M.; Roma, C.; et al. Study of Ras Mutations’ Prognostic Value in Metastatic Colorectal Cancer: STORIA Analysis. Cancers 2020, 12, 1919. [Google Scholar] [CrossRef] [PubMed]
  37. Jones, R.P.; Sutton, P.A.; Evans, J.P.; Clifford, R.; McAvoy, A.; Lewis, J.; Rousseau, A.; Mountford, R.; McWhirter, D.; Malik, H.Z. Specific mutations in KRAS codon 12 are associated with worse overall survival in patients with advanced and recurrent colorectal cancer. Br. J. Cancer 2017, 116, 923–929. [Google Scholar] [CrossRef] [Green Version]
  38. Margonis, G.A.; Kim, Y.; Spolverato, G.; Ejaz, A.; Gupta, R.; Cosgrove, D.; Anders, R.; Karagkounis, G.; Choti, M.A.; Pawlik, T.M. Association Between Specific Mutations in KRAS Codon 12 and Colorectal Liver Metastasis. JAMA Surg. 2015, 150, 722–729. [Google Scholar] [CrossRef] [Green Version]
  39. Koh, Y.H.; Ruiz-Canada, C.; Gorczyca, M.; Budnik, V. The Ras1-mitogen-activated protein kinase signal transduction pathway regulates synaptic plasticity through fasciclin II-mediated cell adhesion. J. Neurosci. 2002, 22, 2496–2504. [Google Scholar] [CrossRef]
  40. Halfar, K.; Rommel, C.; Stocker, H.; Hafen, E. Ras controls growth, survival and differentiation in the Drosophila eye by different thresholds of MAP kinase activity. Development 2001, 128, 1687–1696. [Google Scholar] [CrossRef]
  41. Prober, D.A.; Edgar, B.A. Ras1 promotes cellular growth in the Drosophila wing. Cell 2000, 100, 435–446. [Google Scholar] [CrossRef] [Green Version]
  42. Bergmann, A.; Agapite, J.; McCall, K.; Steller, H. The Drosophila gene hid is a direct molecular target of Ras-dependent survival signaling. Cell 1998, 95, 331–341. [Google Scholar] [CrossRef] [Green Version]
  43. Karim, F.D.; Chang, H.C.; Therrien, M.; Wassarman, D.A.; Laverty, T.; Rubin, G.M. A screen for genes that function downstream of Ras1 during Drosophila eye development. Genetics 1996, 143, 315–329. [Google Scholar] [CrossRef]
  44. Sung, P.J.; Rodrigues, A.B.; Kleinberger, A.; Quatela, S.; Bach, E.A.; Philips, M.R. Cytosolic Ras supports eye development in Drosophila. Mol. Cell Biol. 2010, 30, 5649–5657. [Google Scholar] [CrossRef] [Green Version]
  45. Simon, M.A.; Bowtell, D.D.; Dodson, G.S.; Laverty, T.R.; Rubin, G.M. Ras1 and a putative guanine nucleotide exchange factor perform crucial steps in signaling by the sevenless protein tyrosine kinase. Cell 1991, 67, 701–716. [Google Scholar] [CrossRef]
  46. Wu, Y.; Zhuang, Y.; Han, M.; Xu, T.; Deng, K. Ras promotes cell survival by antagonizing both JNK and Hid signals in the Drosophila eye. BMC Dev. Biol. 2009, 9, 53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Slack, C.; Alic, N.; Foley, A.; Cabecinha, M.; Hoddinott, M.P.; Partridge, L. The Ras-Erk-ETS-Signaling Pathway Is a Drug Target for Longevity. Cell 2015, 162, 72–83. [Google Scholar] [CrossRef] [Green Version]
  48. Prober, D.A.; Edgar, B.A. Interactions between Ras1, dMyc, and dPI3K signaling in the developing Drosophila wing. Genes Dev. 2002, 16, 2286–2299. [Google Scholar] [CrossRef] [Green Version]
  49. Ray, M.; Lakhotia, S.C. Activated Ras/JNK driven Dilp8 in imaginal discs adversely affects organismal homeostasis during early pupal stage in Drosophila, a new checkpoint for development. Dev. Dyn. 2019, 248, 1211–1231. [Google Scholar] [CrossRef] [PubMed]
  50. Karim, F.D.; Rubin, G.M. Ectopic expression of activated Ras1 induces hyperplastic growth and increased cell death in Drosophila imaginal tissues. Development 1998, 125, 1–9. [Google Scholar] [CrossRef]
  51. Hogan, C.; Dupre-Crochet, S.; Norman, M.; Kajita, M.; Zimmermann, C.; Pelling, A.E.; Piddini, E.; Baena-Lopez, L.A.; Vincent, J.P.; Itoh, Y.; et al. Characterization of the interface between normal and transformed epithelial cells. Nat. Cell Biol. 2009, 11, 460–467. [Google Scholar] [CrossRef]
  52. Chi, C.; Zhu, H.; Han, M.; Zhuang, Y.; Wu, X.; Xu, T. Disruption of lysosome function promotes tumor growth and metastasis in Drosophila. J. Biol. Chem. 2010, 285, 21817–21823. [Google Scholar] [CrossRef] [Green Version]
  53. Mishra-Gorur, K.; Li, D.; Ma, X.; Yarman, Y.; Xue, L.; Xu, T. Spz/Toll-6 signal guides organotropic metastasis in Drosophila. Dis. Model. Mech. 2019, 12, dmm039727. [Google Scholar] [CrossRef] [Green Version]
  54. Katheder, N.S.; Khezri, R.; O’Farrell, F.; Schultz, S.W.; Jain, A.; Rahman, M.M.; Schink, K.O.; Theodossiou, T.A.; Johansen, T.; Juhasz, G.; et al. Microenvironmental autophagy promotes tumour growth. Nature 2017, 541, 417–420. [Google Scholar] [CrossRef] [PubMed]
  55. Igaki, T.; Pagliarini, R.A.; Xu, T. Loss of cell polarity drives tumor growth and invasion through JNK activation in Drosophila. Curr. Biol. 2006, 16, 1139–1146. [Google Scholar] [CrossRef] [Green Version]
  56. Uhlirova, M.; Jasper, H.; Bohmann, D. Non-cell-autonomous induction of tissue overgrowth by JNK/Ras cooperation in a Drosophila tumor model. Proc. Natl. Acad. Sci. USA 2005, 102, 13123–13128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Ohsawa, S.; Sugimura, K.; Takino, K.; Xu, T.; Miyawaki, A.; Igaki, T. Elimination of oncogenic neighbors by JNK-mediated engulfment in Drosophila. Dev. Cell 2011, 20, 315–328. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Brumby, A.M.; Goulding, K.R.; Schlosser, T.; Loi, S.; Galea, R.; Khoo, P.; Bolden, J.E.; Aigaki, T.; Humbert, P.O.; Richardson, H.E. Identification of novel Ras-cooperating oncogenes in Drosophila melanogaster: A RhoGEF/Rho-family/JNK pathway is a central driver of tumorigenesis. Genetics 2011, 188, 105–125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. La Marca, J.E.; Richardson, H.E. Two-Faced: Roles of JNK Signalling During Tumourigenesis in the Drosophila Model. Front. Cell Dev. Biol. 2020, 8, 42. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Moreno, E.; Yan, M.; Basler, K. Evolution of TNF signaling mechanisms: JNK-dependent apoptosis triggered by Eiger, the Drosophila homolog of the TNF superfamily. Curr. Biol. 2002, 12, 1263–1268. [Google Scholar] [CrossRef] [Green Version]
  61. Igaki, T.; Kanda, H.; Yamamoto-Goto, Y.; Kanuka, H.; Kuranaga, E.; Aigaki, T.; Miura, M. Eiger, a TNF superfamily ligand that triggers the Drosophila JNK pathway. EMBO J. 2002, 21, 3009–3018. [Google Scholar] [CrossRef] [PubMed]
  62. Kanda, H.; Igaki, T.; Kanuka, H.; Yagi, T.; Miura, M. Wengen, a member of the Drosophila tumor necrosis factor receptor superfamily, is required for Eiger signaling. J. Biol. Chem. 2002, 277, 28372–28375. [Google Scholar] [CrossRef] [Green Version]
  63. Andersen, D.S.; Colombani, J.; Palmerini, V.; Chakrabandhu, K.; Boone, E.; Rothlisberger, M.; Toggweiler, J.; Basler, K.; Mapelli, M.; Hueber, A.O.; et al. The Drosophila TNF receptor Grindelwald couples loss of cell polarity and neoplastic growth. Nature 2015, 522, 482–486. [Google Scholar] [CrossRef]
  64. Palmerini, V.; Monzani, S.; Laurichesse, Q.; Loudhaief, R.; Mari, S.; Cecatiello, V.; Olieric, V.; Pasqualato, S.; Colombani, J.; Andersen, D.S.; et al. Drosophila TNFRs Grindelwald and Wengen bind Eiger with different affinities and promote distinct cellular functions. Nat. Commun. 2021, 12, 2070. [Google Scholar] [CrossRef] [PubMed]
  65. Kanda, H.; Igaki, T.; Okano, H.; Miura, M. Conserved metabolic energy production pathways govern Eiger/TNF-induced nonapoptotic cell death. Proc. Natl. Acad. Sci. USA 2011, 108, 18977–18982. [Google Scholar] [CrossRef] [Green Version]
  66. Igaki, T.; Miura, M. The Drosophila TNF ortholog Eiger: Emerging physiological roles and evolution of the TNF system. Semin. Immunol. 2014, 26, 267–274. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Ma, X.; Chen, Y.; Zhang, S.; Xu, W.; Shao, Y.; Yang, Y.; Li, W.; Li, M.; Xue, L. Rho1-Wnd signaling regulates loss-of-cell polarity-induced cell invasion in Drosophila. Oncogene 2016, 35, 846–855. [Google Scholar] [CrossRef]
  68. Li, M.; Sun, S.; Priest, J.; Bi, X.; Fan, Y. Characterization of TNF-induced cell death in Drosophila reveals caspase- and JNK-dependent necrosis and its role in tumor suppression. Cell Death Dis. 2019, 10, 613. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Leong, G.R.; Goulding, K.R.; Amin, N.; Richardson, H.E.; Brumby, A.M. Scribble mutants promote aPKC and JNK-dependent epithelial neoplasia independently of Crumbs. BMC Biol. 2009, 7, 62. [Google Scholar] [CrossRef] [Green Version]
  70. Kulshammer, E.; Uhlirova, M. The actin cross-linker Filamin/Cheerio mediates tumor malignancy downstream of JNK signaling. J. Cell Sci. 2013, 126, 927–938. [Google Scholar] [CrossRef] [Green Version]
  71. Toggweiler, J.; Willecke, M.; Basler, K. The transcription factor Ets21C drives tumor growth by cooperating with AP-1. Sci. Rep. 2016, 6, 34725. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Kulshammer, E.; Mundorf, J.; Kilinc, M.; Frommolt, P.; Wagle, P.; Uhlirova, M. Interplay among Drosophila transcription factors Ets21c, Fos and Ftz-F1 drives JNK-mediated tumor malignancy. Dis. Model. Mech. 2015, 8, 1279–1293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Doggett, K.; Turkel, N.; Willoughby, L.F.; Ellul, J.; Murray, M.J.; Richardson, H.E.; Brumby, A.M. BTB-Zinc Finger Oncogenes Are Required for Ras and Notch-Driven Tumorigenesis in Drosophila. PLoS ONE 2015, 10, e0132987. [Google Scholar] [CrossRef] [Green Version]
  74. Perez, E.; Lindblad, J.L.; Bergmann, A. Tumor-promoting function of apoptotic caspases by an amplification loop involving ROS, macrophages and JNK in Drosophila. eLife 2017, 6, e26747. [Google Scholar] [CrossRef]
  75. Kurada, P.; White, K. Ras promotes cell survival in Drosophila by downregulating hid expression. Cell 1998, 95, 319–329. [Google Scholar] [CrossRef] [Green Version]
  76. Willecke, M.; Toggweiler, J.; Basler, K. Loss of PI3K blocks cell-cycle progression in a Drosophila tumor model. Oncogene 2011, 30, 4067–4074. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Muzzopappa, M.; Murcia, L.; Milan, M. Feedback amplification loop drives malignant growth in epithelial tissues. Proc. Natl. Acad. Sci. USA 2017, 114, E7291–E7300. [Google Scholar] [CrossRef] [Green Version]
  78. Sun, Y.; Zhang, D.; Guo, X.; Li, W.; Li, C.; Luo, J.; Zhou, M.; Xue, L. MKK3 modulates JNK-dependent cell migration and invasion. Cell Death Dis. 2019, 10, 149. [Google Scholar] [CrossRef]
  79. Ma, X.; Yang, L.; Yang, Y.; Li, M.; Li, W.; Xue, L. dUev1a modulates TNF-JNK mediated tumor progression and cell death in Drosophila. Dev. Biol. 2013, 380, 211–221. [Google Scholar] [CrossRef] [Green Version]
  80. Ma, X.; Shao, Y.; Zheng, H.; Li, M.; Li, W.; Xue, L. Src42A modulates tumor invasion and cell death via Ben/dUev1a-mediated JNK activation in Drosophila. Cell Death Dis. 2013, 4, e864. [Google Scholar] [CrossRef] [Green Version]
  81. Poon, C.L.C.; Brumby, A.M.; Richardson, H.E. Src Cooperates with Oncogenic Ras in Tumourigenesis via the JNK and PI3K Pathways in Drosophila epithelial Tissue. Int. J. Mol. Sci. 2018, 19, 1585. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Li, P.; Ma, Z.; Yu, Y.; Hu, X.; Zhou, Y.; Song, H. FER promotes cell migration via regulating JNK activity. Cell Prolif. 2019, 52, e12656. [Google Scholar] [CrossRef] [PubMed]
  83. Courtial, L.; Picco, V.; Grover, R.; Cormerais, Y.; Rottier, C.; Labbe, A.; Pages, G.; Ferrier-Pages, C. The c-Jun N-terminal kinase prevents oxidative stress induced by UV and thermal stresses in corals and human cells. Sci. Rep. 2017, 7, 45713. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Manent, J.; Banerjee, S.; de Matos Simoes, R.; Zoranovic, T.; Mitsiades, C.; Penninger, J.M.; Simpson, K.J.; Humbert, P.O.; Richardson, H.E. Autophagy suppresses Ras-driven epithelial tumourigenesis by limiting the accumulation of reactive oxygen species. Oncogene 2017, 36, 5658–5660. [Google Scholar] [CrossRef] [Green Version]
  85. Trivedi, S.; Starz-Gaiano, M. Drosophila Jak/STAT Signaling: Regulation and Relevance in Human Cancer and Metastasis. Int. J. Mol. Sci. 2018, 19, 4056. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Wu, M.; Pastor-Pareja, J.C.; Xu, T. Interaction between Ras(V12) and scribbled clones induces tumour growth and invasion. Nature 2010, 463, 545–548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Bunker, B.D.; Nellimoottil, T.T.; Boileau, R.M.; Classen, A.K.; Bilder, D. The transcriptional response to tumorigenic polarity loss in Drosophila. eLife 2015, 4, e03189. [Google Scholar] [CrossRef]
  88. He, Z.; Li, R.; Jiang, H. Mutations and Copy Number Abnormalities of Hippo Pathway Components in Human Cancers. Front. Cell Dev. Biol. 2021, 9, 661718. [Google Scholar] [CrossRef]
  89. Snigdha, K.; Gangwani, K.S.; Lapalikar, G.V.; Singh, A.; Kango-Singh, M. Hippo Signaling in Cancer: Lessons From Drosophila Models. Front. Cell Dev. Biol. 2019, 7, 85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Doggett, K.; Grusche, F.A.; Richardson, H.E.; Brumby, A.M. Loss of the Drosophila cell polarity regulator Scribbled promotes epithelial tissue overgrowth and cooperation with oncogenic Ras-Raf through impaired Hippo pathway signaling. BMC Dev. Biol. 2011, 11, 57. [Google Scholar] [CrossRef] [Green Version]
  91. Menendez, J.; Perez-Garijo, A.; Calleja, M.; Morata, G. A tumor-suppressing mechanism in Drosophila involving cell competition and the Hippo pathway. Proc. Natl. Acad. Sci. USA 2010, 107, 14651–14656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Chen, C.L.; Schroeder, M.C.; Kango-Singh, M.; Tao, C.; Halder, G. Tumor suppression by cell competition through regulation of the Hippo pathway. Proc. Natl. Acad. Sci. USA 2012, 109, 484–489. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Sun, G.; Irvine, K.D. Regulation of Hippo signaling by Jun kinase signaling during compensatory cell proliferation and regeneration, and in neoplastic tumors. Dev. Biol. 2011, 350, 139–151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Snigdha, K.; Singh, A.; Kango-Singh, M. Yorkie-Cactus (IkappaBalpha)-JNK axis promotes tumor growth and progression in Drosophila. Oncogene 2021, 40, 4124–4136. [Google Scholar] [CrossRef] [PubMed]
  95. Atkins, M.; Potier, D.; Romanelli, L.; Jacobs, J.; Mach, J.; Hamaratoglu, F.; Aerts, S.; Halder, G. An Ectopic Network of Transcription Factors Regulated by Hippo Signaling Drives Growth and Invasion of a Malignant Tumor Model. Curr. Biol. 2016, 26, 2101–2113. [Google Scholar] [CrossRef] [Green Version]
  96. Wang, X.; Liang, H.; Xu, W.; Ma, X. Wallenda-Nmo Axis Regulates Growth via Hippo Signaling. Front. Cell Dev. Biol. 2021, 9, 658288. [Google Scholar] [CrossRef]
  97. Ohsawa, S.; Sato, Y.; Enomoto, M.; Nakamura, M.; Betsumiya, A.; Igaki, T. Mitochondrial defect drives non-autonomous tumour progression through Hippo signalling in Drosophila. Nature 2012, 490, 547–551. [Google Scholar] [CrossRef]
  98. Kong, D.; Lu, J.Y.; Li, X.; Zhao, S.; Xu, W.; Fang, J.; Wang, X.; Ma, X. Misshapen Disruption Cooperates with Ras(V12) to Drive Tumorigenesis. Cells 2021, 10, 894. [Google Scholar] [CrossRef]
  99. Enomoto, M.; Kizawa, D.; Ohsawa, S.; Igaki, T. JNK signaling is converted from anti- to pro-tumor pathway by Ras-mediated switch of Warts activity. Dev. Biol. 2015, 403, 162–171. [Google Scholar] [CrossRef] [Green Version]
  100. Davie, K.; Jacobs, J.; Atkins, M.; Potier, D.; Christiaens, V.; Halder, G.; Aerts, S. Discovery of transcription factors and regulatory regions driving in vivo tumor development by ATAC-seq and FAIRE-seq open chromatin profiling. PLoS Genet. 2015, 11, e1004994. [Google Scholar] [CrossRef] [Green Version]
  101. Ferrari, A.J.; Drapkin, R.; Gogna, R. Cell Fitness: More Than Push-Ups. Int. J. Mol. Sci. 2021, 22, 518. [Google Scholar] [CrossRef]
  102. Simpson, P. Parameters of cell competition in the compartments of the wing disc of Drosophila. Dev. Biol. 1979, 69, 182–193. [Google Scholar] [CrossRef]
  103. Bilder, D.; Li, M.; Perrimon, N. Cooperative regulation of cell polarity and growth by Drosophila tumor suppressors. Science 2000, 289, 113–116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Katsukawa, M.; Ohsawa, S.; Zhang, L.; Yan, Y.; Igaki, T. Serpin Facilitates Tumor-Suppressive Cell Competition by Blocking Toll-Mediated Yki Activation in Drosophila. Curr. Biol. 2018, 28, 1756–1767.e1756. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Ziosi, M.; Baena-Lopez, L.A.; Grifoni, D.; Froldi, F.; Pession, A.; Garoia, F.; Trotta, V.; Bellosta, P.; Cavicchi, S.; Pession, A. dMyc functions downstream of Yorkie to promote the supercompetitive behavior of hippo pathway mutant cells. PLoS Genet. 2010, 6, e1001140. [Google Scholar] [CrossRef] [Green Version]
  106. Yamamoto, M.; Ohsawa, S.; Kunimasa, K.; Igaki, T. The ligand Sas and its receptor PTP10D drive tumour-suppressive cell competition. Nature 2017, 542, 246–250. [Google Scholar] [CrossRef] [PubMed]
  107. Lee, H.K.; Cording, A.; Vielmetter, J.; Zinn, K. Interactions between a receptor tyrosine phosphatase and a cell surface ligand regulate axon guidance and glial-neuronal communication. Neuron 2013, 78, 813–826. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Tarcic, G.; Boguslavsky, S.K.; Wakim, J.; Kiuchi, T.; Liu, A.; Reinitz, F.; Nathanson, D.; Takahashi, T.; Mischel, P.S.; Ng, T.; et al. An unbiased screen identifies DEP-1 tumor suppressor as a phosphatase controlling EGFR endocytosis. Curr. Biol. 2009, 19, 1788–1798. [Google Scholar] [CrossRef] [Green Version]
  109. Florey, O.; Krajcovic, M.; Sun, Q.; Overholtzer, M. Entosis. Curr. Biol. 2010, 20, R88–R89. [Google Scholar] [CrossRef] [Green Version]
  110. Geisbrecht, E.R.; Haralalka, S.; Swanson, S.K.; Florens, L.; Washburn, M.P.; Abmayr, S.M. Drosophila ELMO/CED-12 interacts with Myoblast city to direct myoblast fusion and ommatidial organization. Dev. Biol. 2008, 314, 137–149. [Google Scholar] [CrossRef] [Green Version]
  111. Li, W.; Baker, N.E. Engulfment is required for cell competition. Cell 2007, 129, 1215–1225. [Google Scholar] [CrossRef] [Green Version]
  112. Valon, L.; Davidovic, A.; Levillayer, F.; Villars, A.; Chouly, M.; Cerqueira-Campos, F.; Levayer, R. Robustness of epithelial sealing is an emerging property of local ERK feedback driven by cell elimination. Dev. Cell 2021, 56, 1700–1711.e8. [Google Scholar] [CrossRef]
  113. Diwanji, N.; Bergmann, A. Two Sides of the Same Coin–Compensatory Proliferation in Regeneration and Cancer. Adv. Exp. Med. Biol. 2019, 1167, 65–85. [Google Scholar] [CrossRef]
  114. Fogarty, C.E.; Diwanji, N.; Lindblad, J.L.; Tare, M.; Amcheslavsky, A.; Makhijani, K.; Bruckner, K.; Fan, Y.; Bergmann, A. Extracellular Reactive Oxygen Species Drive Apoptosis-Induced Proliferation via Drosophila Macrophages. Curr. Biol. 2016, 26, 575–584. [Google Scholar] [CrossRef] [Green Version]
  115. Khezri, R.; Holland, P.; Schoborg, T.A.; Abramovich, I.; Takats, S.; Dillard, C.; Jain, A.; O’Farrell, F.; Schultz, S.W.; Hagopian, W.M.; et al. Host autophagy mediates organ wasting and nutrient mobilization for tumor growth. EMBO J. 2021, e107336. [Google Scholar] [CrossRef] [PubMed]
  116. Sousa, C.M.; Biancur, D.E.; Wang, X.; Halbrook, C.J.; Sherman, M.H.; Zhang, L.; Kremer, D.; Hwang, R.F.; Witkiewicz, A.K.; Ying, H.; et al. Pancreatic stellate cells support tumour metabolism through autophagic alanine secretion. Nature 2016, 536, 479–483. [Google Scholar] [CrossRef] [Green Version]
  117. Khezri, R.; Rusten, T.E. Autophagy and Tumorigenesis in Drosophila. Adv. Exp. Med. Biol. 2019, 1167, 113–127. [Google Scholar] [CrossRef]
  118. Tamori, Y.; Suzuki, E.; Deng, W.M. Epithelial Tumors Originate in Tumor Hotspots, a Tissue-Intrinsic Microenvironment. PLoS Biol. 2016, 14, e1002537. [Google Scholar] [CrossRef]
  119. Yamanaka, N.; Rewitz, K.F.; O’Connor, M.B. Ecdysone control of developmental transitions: Lessons from Drosophila research. Annu. Rev. Entomol. 2013, 58, 497–516. [Google Scholar] [CrossRef] [Green Version]
  120. Garelli, A.; Gontijo, A.M.; Miguela, V.; Caparros, E.; Dominguez, M. Imaginal discs secrete insulin-like peptide 8 to mediate plasticity of growth and maturation. Science 2012, 336, 579–582. [Google Scholar] [CrossRef]
  121. Colombani, J.; Andersen, D.S.; Leopold, P. Secreted peptide Dilp8 coordinates Drosophila tissue growth with developmental timing. Science 2012, 336, 582–585. [Google Scholar] [CrossRef]
  122. Jaszczak, J.S.; Wolpe, J.B.; Bhandari, R.; Jaszczak, R.G.; Halme, A. Growth Coordination During Drosophila melanogaster Imaginal Disc Regeneration Is Mediated by Signaling Through the Relaxin Receptor Lgr3 in the Prothoracic Gland. Genetics 2016, 204, 703–709. [Google Scholar] [CrossRef] [Green Version]
  123. Vallejo, D.M.; Juarez-Carreno, S.; Bolivar, J.; Morante, J.; Dominguez, M. A brain circuit that synchronizes growth and maturation revealed through Dilp8 binding to Lgr3. Science 2015, 350, aac6767. [Google Scholar] [CrossRef] [Green Version]
  124. Garelli, A.; Heredia, F.; Casimiro, A.P.; Macedo, A.; Nunes, C.; Garcez, M.; Dias, A.R.M.; Volonte, Y.A.; Uhlmann, T.; Caparros, E.; et al. Dilp8 requires the neuronal relaxin receptor Lgr3 to couple growth to developmental timing. Nat. Commun. 2015, 6, 8732. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Colombani, J.; Andersen, D.S.; Boulan, L.; Boone, E.; Romero, N.; Virolle, V.; Texada, M.; Leopold, P. Drosophila Lgr3 Couples Organ Growth with Maturation and Ensures Developmental Stability. Curr. Biol. 2015, 25, 2723–2729. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Figueroa-Clarevega, A.; Bilder, D. Malignant Drosophila tumors interrupt insulin signaling to induce cachexia-like wasting. Dev. Cell 2015, 33, 47–55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Boone, E.; Colombani, J.; Andersen, D.S.; Leopold, P. The Hippo signalling pathway coordinates organ growth and limits developmental variability by controlling dilp8 expression. Nat. Commun. 2016, 7, 13505. [Google Scholar] [CrossRef]
  128. Romao, D.; Muzzopappa, M.; Barrio, L.; Milan, M. The Upd3 cytokine couples inflammation to maturation defects in Drosophila. Curr. Biol. 2021, 31, 1780–1787. [Google Scholar] [CrossRef]
  129. Banerjee, U.; Girard, J.R.; Goins, L.M.; Spratford, C.M. Drosophila as a Genetic Model for Hematopoiesis. Genetics 2019, 211, 367–417. [Google Scholar] [CrossRef] [Green Version]
  130. Parisi, F.; Stefanatos, R.K.; Strathdee, K.; Yu, Y.; Vidal, M. Transformed epithelia trigger non-tissue-autonomous tumor suppressor response by adipocytes via activation of Toll and Eiger/TNF signaling. Cell Rep. 2014, 6, 855–867. [Google Scholar] [CrossRef] [Green Version]
  131. Rossi, M.J. Editorial Commentary: Iliotibial Band Allograft Shows Promise for Arthroscopic Hip Labral Reconstruction. Arthroscopy 2016, 32, 33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Parvy, J.P.; Yu, Y.; Dostalova, A.; Kondo, S.; Kurjan, A.; Bulet, P.; Lemaitre, B.; Vidal, M.; Cordero, J.B. The antimicrobial peptide defensin cooperates with tumour necrosis factor to drive tumour cell death in Drosophila. eLife 2019, 8, e45061. [Google Scholar] [CrossRef] [PubMed]
  133. Diwanji, N.; Bergmann, A. Basement membrane damage by ROS- and JNK-mediated Mmp2 activation drives macrophage recruitment to overgrown tissue. Nat. Commun. 2020, 11, 3631. [Google Scholar] [CrossRef]
  134. Jacqueline, C.; Parvy, J.P.; Rollin, M.L.; Faugere, D.; Renaud, F.; Misse, D.; Thomas, F.; Roche, B. The role of innate immunity in the protection conferred by a bacterial infection against cancer: Study of an invertebrate model. Sci. Rep. 2020, 10, 10106. [Google Scholar] [CrossRef] [PubMed]
  135. Song, W.; Kir, S.; Hong, S.; Hu, Y.; Wang, X.; Binari, R.; Tang, H.W.; Chung, V.; Banks, A.S.; Spiegelman, B.; et al. Tumor-Derived Ligands Trigger Tumor Growth and Host Wasting via Differential MEK Activation. Dev. Cell 2019, 48, 277–286.e276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Kwon, Y.; Song, W.; Droujinine, I.A.; Hu, Y.; Asara, J.M.; Perrimon, N. Systemic organ wasting induced by localized expression of the secreted insulin/IGF antagonist ImpL2. Dev. Cell 2015, 33, 36–46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Khormali, K.; Mizwari, Z.M.; Masoumeh Ghoreishi, S.; Mortazavi-Derazkola, S.; Khezri, B. Novel Dy2O3/ZnO-Au ternary nanocomposites: Green synthesis using pomegranate fruit extract, characterization and their photocatalytic and antibacterial properties. Bioorg. Chem. 2021, 115, 105204. [Google Scholar] [CrossRef]
  138. Newton, H.; Wang, Y.F.; Camplese, L.; Mokochinski, J.B.; Kramer, H.B.; Brown, A.E.X.; Fets, L.; Hirabayashi, S. Systemic muscle wasting and coordinated tumour response drive tumourigenesis. Nat. Commun. 2020, 11, 4653. [Google Scholar] [CrossRef]
  139. Saavedra, P.; Perrimon, N. Drosophila as a Model for Tumor-Induced Organ Wasting. Adv. Exp. Med. Biol. 2019, 1167, 191–205. [Google Scholar] [CrossRef]
  140. Holland, P.; Hagopian, W.M.; Jahren, A.H.; Rusten, T.E. Natural abundance isotope ratios to differentiate sources of carbon used during tumor growth in vivo. BMC Biol. 2021, 19, 85. [Google Scholar] [CrossRef]
  141. Hamaratoglu, F.; Atkins, M. Rounding up the Usual Suspects: Assessing Yorkie, AP-1, and Stat Coactivation in Tumorigenesis. Int. J. Mol. Sci. 2020, 21, 4580. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The cooperation between the oncogene RasV12 and loss of the tumor suppressor scribble drives the malignant transformation of eye disc epithelial cells. (A) Representation of a larva with most of its organs. An eye-antennal disc is composed of two main parts. The anterior part is the antennal part (white and grey), which will transform into the antenna and part of the head during metamorphosis. The posterior part is the eye part which generates the photoneurons (purple) in a synchronous manner after the morphogenetic furrow (purple line). (B) scrib−/− clones (green) form benign tumors characterized by a loss of cell polarity, resistance to differentiation into neurons (accompanied by a distortion of the morphogenetic furrow), overproliferation, and basement membrane degradation. Ultimately, they are eliminated through apoptosis. (C) RasV12 clones (green) form hyperplastic tumors (defined as tumors that overgrow without disrupting the tissue architecture and differentiation). They are characterized by overproliferation and cell growth. Because neuronal differentiation is not impaired within RasV12 clones, transformed cells differentiate synchronously to their wt neighbors (purple cells within the clones) at the morphogenetic furrow (thick purple line) and form long neuronal axons (purple lines) that extend through the posterior part of the eye disc in order to form the optic nerve. (D) RasV12; scrib−/− clones form neoplastic tumors (defined as tumors that disrupt the tissue morphology and impair differentiation). They are characterized by a loss of cell polarity, resistance to differentiation, overproliferation, cell growth, basement membrane degradation, and invasion. RasV12; scrib−/− tumors trigger a developmental delay, postponing pupation.
Figure 1. The cooperation between the oncogene RasV12 and loss of the tumor suppressor scribble drives the malignant transformation of eye disc epithelial cells. (A) Representation of a larva with most of its organs. An eye-antennal disc is composed of two main parts. The anterior part is the antennal part (white and grey), which will transform into the antenna and part of the head during metamorphosis. The posterior part is the eye part which generates the photoneurons (purple) in a synchronous manner after the morphogenetic furrow (purple line). (B) scrib−/− clones (green) form benign tumors characterized by a loss of cell polarity, resistance to differentiation into neurons (accompanied by a distortion of the morphogenetic furrow), overproliferation, and basement membrane degradation. Ultimately, they are eliminated through apoptosis. (C) RasV12 clones (green) form hyperplastic tumors (defined as tumors that overgrow without disrupting the tissue architecture and differentiation). They are characterized by overproliferation and cell growth. Because neuronal differentiation is not impaired within RasV12 clones, transformed cells differentiate synchronously to their wt neighbors (purple cells within the clones) at the morphogenetic furrow (thick purple line) and form long neuronal axons (purple lines) that extend through the posterior part of the eye disc in order to form the optic nerve. (D) RasV12; scrib−/− clones form neoplastic tumors (defined as tumors that disrupt the tissue morphology and impair differentiation). They are characterized by a loss of cell polarity, resistance to differentiation, overproliferation, cell growth, basement membrane degradation, and invasion. RasV12; scrib−/− tumors trigger a developmental delay, postponing pupation.
Ijms 22 08873 g001
Figure 2. Cooperative oncogenesis results from the atypical interplay of various signaling pathways that display several tumor cancer hallmarks. (A) In scrib−/− benign tumors, the JNK pathway (green) has an antitumorigenic function. It is activated by the TNF Eiger, which binds to the TNFR Wengen (Wgn). Elevated endocytosis levels contribute to activating the Eiger/Wgn complex that activates a phosphorylation event cascade involving successively the JNKKKK, JNKKK, and JNKK. Ultimately, the JNK Basket is phosphorylated and activates its downstream AP-1 TFs effectors. AP-1 TFs stimulates the expression of the proapoptotic factors rpr and hid, which execute apoptosis a well as paxillin and mmp1, which leads to basement membrane degradation. Simultaneously, the loss of cell polarity initiated by the loss of function of scribble provokes aPKC (blue) signaling impairment, which results in the activation of the SWH target Yki (orange). Yki together with Sd activates the expression of cyclinE and causes overproliferation. Finally, DE-cadherin expression decreases, contributing further to the loss of cell polarity phenotype. (B) In RasV12 hyperplastic tumors, the Scrib (grey) and aPKC (blue) apical complexes usually are localized close to the adherens junction and in the apical part, respectively. The aberrantly active RasV12 stimulates the downstream Raf/Mek/MAPK and PI3K/AKT pathways (purple) responsible for cyclinE (overproliferation) and dmyc (cell growth) overexpression, respectively. (C) In the RasV12; scrib−/− neoplastic tumors, JNK is protumorigenic. The upstream JNK signaling is not fully explored in the eye-antennal disc tumors yet. It seems to involve Eiger binding to Grnd mainly (and Wgn), which further activates the JNKKKK Msn, the JNKKK Tak1, the JNKK Hep, and Bsk. A parallel Src-dUev1/Ben/dTraf2 axis elevates the activity of Hep upstream of Bsk. The novel JNKK MMK3 and the nonreceptor tyrosine kinase FER might also contribute to Bsk activity. Bsk triggers the activation of the TFs Fos, Ets21c, and Ftfz1, which upregulate the expression of the BTBZF chinmo and abrupt (resistance to neuron differentiation), mmp1 and cher (basement membrane degradation and invasion), as well as dilp8 and upd (developmental delay). Whether they also elevate hid transcription is unknown. In parallel, lowering of SWH signaling leads to the activation of Yki (orange). Together with Sd, Yki activates the expression of dmyc (cell growth), cyclinE (proliferation), dilp8 and upd (developmental delay, not demonstrated) and of some other signaling components such as upd and stat (JAK-STAT pathway) as well as fos and ftzf1 (JNK pathway). RasV12 activates the Raf/Mek/MAPK and PI3K/AKT pathways, contributing to the expression of dmyc (cell growth) and cyclinE (proliferation). It also inhibits hid expression and Hid activity, protecting RasV12; scrib−/− cancer cells from apoptosis. Finally, tumor-secreted Upds activate the JAK-STAT pathway in an autocrine manner and contribute to ROS production and tumor growth through an unknown mechanism.
Figure 2. Cooperative oncogenesis results from the atypical interplay of various signaling pathways that display several tumor cancer hallmarks. (A) In scrib−/− benign tumors, the JNK pathway (green) has an antitumorigenic function. It is activated by the TNF Eiger, which binds to the TNFR Wengen (Wgn). Elevated endocytosis levels contribute to activating the Eiger/Wgn complex that activates a phosphorylation event cascade involving successively the JNKKKK, JNKKK, and JNKK. Ultimately, the JNK Basket is phosphorylated and activates its downstream AP-1 TFs effectors. AP-1 TFs stimulates the expression of the proapoptotic factors rpr and hid, which execute apoptosis a well as paxillin and mmp1, which leads to basement membrane degradation. Simultaneously, the loss of cell polarity initiated by the loss of function of scribble provokes aPKC (blue) signaling impairment, which results in the activation of the SWH target Yki (orange). Yki together with Sd activates the expression of cyclinE and causes overproliferation. Finally, DE-cadherin expression decreases, contributing further to the loss of cell polarity phenotype. (B) In RasV12 hyperplastic tumors, the Scrib (grey) and aPKC (blue) apical complexes usually are localized close to the adherens junction and in the apical part, respectively. The aberrantly active RasV12 stimulates the downstream Raf/Mek/MAPK and PI3K/AKT pathways (purple) responsible for cyclinE (overproliferation) and dmyc (cell growth) overexpression, respectively. (C) In the RasV12; scrib−/− neoplastic tumors, JNK is protumorigenic. The upstream JNK signaling is not fully explored in the eye-antennal disc tumors yet. It seems to involve Eiger binding to Grnd mainly (and Wgn), which further activates the JNKKKK Msn, the JNKKK Tak1, the JNKK Hep, and Bsk. A parallel Src-dUev1/Ben/dTraf2 axis elevates the activity of Hep upstream of Bsk. The novel JNKK MMK3 and the nonreceptor tyrosine kinase FER might also contribute to Bsk activity. Bsk triggers the activation of the TFs Fos, Ets21c, and Ftfz1, which upregulate the expression of the BTBZF chinmo and abrupt (resistance to neuron differentiation), mmp1 and cher (basement membrane degradation and invasion), as well as dilp8 and upd (developmental delay). Whether they also elevate hid transcription is unknown. In parallel, lowering of SWH signaling leads to the activation of Yki (orange). Together with Sd, Yki activates the expression of dmyc (cell growth), cyclinE (proliferation), dilp8 and upd (developmental delay, not demonstrated) and of some other signaling components such as upd and stat (JAK-STAT pathway) as well as fos and ftzf1 (JNK pathway). RasV12 activates the Raf/Mek/MAPK and PI3K/AKT pathways, contributing to the expression of dmyc (cell growth) and cyclinE (proliferation). It also inhibits hid expression and Hid activity, protecting RasV12; scrib−/− cancer cells from apoptosis. Finally, tumor-secreted Upds activate the JAK-STAT pathway in an autocrine manner and contribute to ROS production and tumor growth through an unknown mechanism.
Ijms 22 08873 g002
Figure 3. Local microenvironment interactions can act as suppressors or be exploited for tumor growth. (A) Anti-tumorigenic role of the microenvironment. scrib−/− tumors are actively eliminated by the neighboring cells through cell competition. Several mechanisms are put in place by the winner cells, the microenvironment, to restrict the growth and remove the loser cells, the scrib−/− tumors. (a) The expression of an inhibitor of the toll ligand Spz, Spn5, is critical in the microenvironment to maintain the loser status of scrib−/− tumors. Upon non-autonomous loss of spn5, Toll signaling, together with JNK and high F-actin levels, can elevate yki activity, an inducer of the winner status (shown in red). (b) Upon tumorigenesis, wt cells communicate with the tumor cells through the Sas-Ptp10D signaling axis to inhibit Egfr-Ras signaling, a pathway that can potently cooperate with polarity deficiency for malignancy. (c) Activation of JNK in the wt cells activates an engulfment program through PVR-ELMO/MBC which leads to the entosis of the scrib−/− tumors and their death. (B) Pro-tumorigenic role of the microenvironment. In a cooperative RasV12; scrib−/− scenario, the microenvironment is exploited to fuel tumor growth. (a) Caspase apoptotic induction is inhibited by RasV12; however, apoptotic-independent functions still remain active. RasV12; scrib−/− tumors exploit the mechanism behind apoptosis induced proliferation. Caspases, and potentially impaired mitochondria, can induce high levels of ROS, which triggers hemocyte recruitment. This mounts an inflammatory response that leads to higher JNK levels. (b) NAA is induced in the wt cells surrounding RasV12; scrib−/− tumors, through tumor autonomous JNK and the autocrine activation of JAK-STAT through Upd and Dome. Downregulation of the transporter slif in RasV12; dlg−/− showed that these tumors are dependent on exogenous nutrient sources, potentially released from the neighboring wt cells through NAA. (c) Areas that are nurturing or hindering for tumor growth, termed tumor “hotspots” and “coldspots”, respectively, have been identified in the wing disc. The tumor hotspot, which is in the wing disc hinge, was found to have high levels of JAK-STAT, a pathway critical for the RasV12; scrib−/− tumor growth, and apical extrusion which promotes tumor cell survival. (d) The RasV12; scrib−/− cooperative scenario is not restricted to an intraclonal setting. Neighboring scrib−/− clones to RasV12 tumors has been found to be sufficient to elevate JNK and JAK-STAT in RasV12 tumors, and to trigger their large overgrowth and invasion.
Figure 3. Local microenvironment interactions can act as suppressors or be exploited for tumor growth. (A) Anti-tumorigenic role of the microenvironment. scrib−/− tumors are actively eliminated by the neighboring cells through cell competition. Several mechanisms are put in place by the winner cells, the microenvironment, to restrict the growth and remove the loser cells, the scrib−/− tumors. (a) The expression of an inhibitor of the toll ligand Spz, Spn5, is critical in the microenvironment to maintain the loser status of scrib−/− tumors. Upon non-autonomous loss of spn5, Toll signaling, together with JNK and high F-actin levels, can elevate yki activity, an inducer of the winner status (shown in red). (b) Upon tumorigenesis, wt cells communicate with the tumor cells through the Sas-Ptp10D signaling axis to inhibit Egfr-Ras signaling, a pathway that can potently cooperate with polarity deficiency for malignancy. (c) Activation of JNK in the wt cells activates an engulfment program through PVR-ELMO/MBC which leads to the entosis of the scrib−/− tumors and their death. (B) Pro-tumorigenic role of the microenvironment. In a cooperative RasV12; scrib−/− scenario, the microenvironment is exploited to fuel tumor growth. (a) Caspase apoptotic induction is inhibited by RasV12; however, apoptotic-independent functions still remain active. RasV12; scrib−/− tumors exploit the mechanism behind apoptosis induced proliferation. Caspases, and potentially impaired mitochondria, can induce high levels of ROS, which triggers hemocyte recruitment. This mounts an inflammatory response that leads to higher JNK levels. (b) NAA is induced in the wt cells surrounding RasV12; scrib−/− tumors, through tumor autonomous JNK and the autocrine activation of JAK-STAT through Upd and Dome. Downregulation of the transporter slif in RasV12; dlg−/− showed that these tumors are dependent on exogenous nutrient sources, potentially released from the neighboring wt cells through NAA. (c) Areas that are nurturing or hindering for tumor growth, termed tumor “hotspots” and “coldspots”, respectively, have been identified in the wing disc. The tumor hotspot, which is in the wing disc hinge, was found to have high levels of JAK-STAT, a pathway critical for the RasV12; scrib−/− tumor growth, and apical extrusion which promotes tumor cell survival. (d) The RasV12; scrib−/− cooperative scenario is not restricted to an intraclonal setting. Neighboring scrib−/− clones to RasV12 tumors has been found to be sufficient to elevate JNK and JAK-STAT in RasV12 tumors, and to trigger their large overgrowth and invasion.
Ijms 22 08873 g003
Figure 4. RasV12; scrib−/− eye antennal disc tumors alter distant organ functions. Invasion: RasV12; scrib−/− tumors (green) invade adjacent organs such as the central nervous system (CNS) and the leg discs (LD). Invasion is guided by the interaction between the Toll6 receptor at the surface of invasive tumor cells and its ligand Spz, secreted by a subset of organs prone to secondary tumor formation. Developmental delay: It relies on the decrease of Ecdysone production by the prothoracic gland (PG). The tumor cells secrete Dilp8 and Upd3, which are both known for their ability to inhibit Ecdysone production. Inflammation: Tumor cells (green) have elevated ROS levels, attracting plasmatocytes (purple) to the basement membrane breach. These plasmatocytes secrete Eiger and contribute to mmp1 expression in the tumor cells and subsequently to their ability to escape the primary tumor, arguing for a pro tumorigenic function. Together with the fat body, they constitute the potential source of the AMP Drosomycin that is found elevated in the larval hemolymph. Drosomycin (Drs) triggers cell death in the tumor, arguing for a potential anti-tumorigenic role of the plasmatocytes. Wasting: Tumor cells secrete ImpL2 (and potentially Pvf and Bnl), which is thought to trigger a wasting phenotype in the larva characterized by the morphological deterioration of the energy-storing organs: the muscles (shrinkage and split) and the fat body (pearl-like phenotype, lipid droplet aggregation). This is accompanied by elevation of autophagy in these organs as well as elevated sugar and amino acid levels in the hemolymph. In its late stage, the tumor building blocks derive from host to maintain its growth.
Figure 4. RasV12; scrib−/− eye antennal disc tumors alter distant organ functions. Invasion: RasV12; scrib−/− tumors (green) invade adjacent organs such as the central nervous system (CNS) and the leg discs (LD). Invasion is guided by the interaction between the Toll6 receptor at the surface of invasive tumor cells and its ligand Spz, secreted by a subset of organs prone to secondary tumor formation. Developmental delay: It relies on the decrease of Ecdysone production by the prothoracic gland (PG). The tumor cells secrete Dilp8 and Upd3, which are both known for their ability to inhibit Ecdysone production. Inflammation: Tumor cells (green) have elevated ROS levels, attracting plasmatocytes (purple) to the basement membrane breach. These plasmatocytes secrete Eiger and contribute to mmp1 expression in the tumor cells and subsequently to their ability to escape the primary tumor, arguing for a pro tumorigenic function. Together with the fat body, they constitute the potential source of the AMP Drosomycin that is found elevated in the larval hemolymph. Drosomycin (Drs) triggers cell death in the tumor, arguing for a potential anti-tumorigenic role of the plasmatocytes. Wasting: Tumor cells secrete ImpL2 (and potentially Pvf and Bnl), which is thought to trigger a wasting phenotype in the larva characterized by the morphological deterioration of the energy-storing organs: the muscles (shrinkage and split) and the fat body (pearl-like phenotype, lipid droplet aggregation). This is accompanied by elevation of autophagy in these organs as well as elevated sugar and amino acid levels in the hemolymph. In its late stage, the tumor building blocks derive from host to maintain its growth.
Ijms 22 08873 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Dillard, C.; Reis, J.G.T.; Rusten, T.E. RasV12; scrib−/− Tumors: A Cooperative Oncogenesis Model Fueled by Tumor/Host Interactions. Int. J. Mol. Sci. 2021, 22, 8873. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22168873

AMA Style

Dillard C, Reis JGT, Rusten TE. RasV12; scrib−/− Tumors: A Cooperative Oncogenesis Model Fueled by Tumor/Host Interactions. International Journal of Molecular Sciences. 2021; 22(16):8873. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22168873

Chicago/Turabian Style

Dillard, Caroline, José Gerardo Teles Reis, and Tor Erik Rusten. 2021. "RasV12; scrib−/− Tumors: A Cooperative Oncogenesis Model Fueled by Tumor/Host Interactions" International Journal of Molecular Sciences 22, no. 16: 8873. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms22168873

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop