Next Article in Journal
Interaction and Metabolic Pathways: Elucidating the Role of Gut Microbiota in Gestational Diabetes Mellitus Pathogenesis
Previous Article in Journal
The Urinary Metabolome of Newborns with Perinatal Complications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Obesity, Dietary Fats, and Gastrointestinal Cancer Risk-Potential Mechanisms Relating to Lipid Metabolism and Inflammation

by
Kathleen A. J. Mitchelson
1,
Fiona O’Connell
2,
Jacintha O’Sullivan
2 and
Helen M. Roche
1,3,*
1
Nutrigenomics Research Group, UCD Conway Institute, UCD Institute of Food and Health, and School of Public Health, Physiotherapy and Sports Science, University College Dublin, D04 H1W8 Dublin, Ireland
2
Department of Surgery, Trinity St. James’s Cancer Institute and Trinity Translational Medicine Institute, St. James’s Hospital and Trinity College Dublin, D08 W9RT Dublin, Ireland
3
Institute for Global Food Security, School of Biological Sciences, Queens University Belfast, Belfast BT9 5DL, UK
*
Author to whom correspondence should be addressed.
Submission received: 28 November 2023 / Revised: 28 December 2023 / Accepted: 29 December 2023 / Published: 10 January 2024
(This article belongs to the Section Nutrition and Metabolism)

Abstract

:
Obesity is a major driving factor in the incidence, progression, and poor treatment response in gastrointestinal cancers. Herein, we conducted a comprehensive analysis of the impact of obesity and its resulting metabolic perturbations across four gastrointestinal cancer types, namely, oesophageal, gastric, liver, and colorectal cancer. Importantly, not all obese phenotypes are equal. Obese adipose tissue heterogeneity depends on the location, structure, cellular profile (including resident immune cell populations), and dietary fatty acid intake. We discuss whether adipose heterogeneity impacts the tumorigenic environment. Dietary fat quality, in particular saturated fatty acids, promotes a hypertrophic, pro-inflammatory adipose profile, in contrast to monounsaturated fatty acids, resulting in a hyperplastic, less inflammatory adipose phenotype. The purpose of this review is to examine the impact of obesity, including dietary fat quality, on adipose tissue biology and oncogenesis, specifically focusing on lipid metabolism and inflammatory mechanisms. This is achieved with a particular focus on gastrointestinal cancers as exemplar models of obesity-associated cancers.

1. Overview: Obesity, Dietary Fats, and Cancer

Global obesity rates have tripled since the 1970s. The causal relationship between obesity and several metabolic co-morbidities, including insulin resistance (IR), type 2 diabetes (T2D), and cardiovascular disease (CVD), is well characterised [1]. Indeed, obesity has emerged as a major determinant of some cancers, overtaking smoking as a leading cause [2]. There are 14 types of cancer linked to obesity, including gastrointestinal cancers such as oesophageal, gastric, liver, and colorectal cancer [3]. Importantly, obesity is a very heterogeneous condition, the impact of which, in terms of the associated metabolic and inflammatory phenotypes, differs greatly between individuals. For example, for an equivalent body weight or body mass index (BMI), some people are profoundly insulin resistant at a given body weight/adiposity, while others remain insulin sensitive [4]. Diet may be one of the driving factors contributing to these differences in metabolic phenotypes, disease, and subsequent cancer incidence, progression, and therapeutic response. Indeed, diet is an important modulator of ‘metabolic inflammation’, a cellular phenomenon wherein the metabolic configuration of an immune cell determines and drives the nature of the inflammatory response [5]. Work to date shows that saturated fatty acids (SFA) and monounsaturated fatty acids (MUFA) have differential effects on metabolism and inflammation, and thus potentially directly impact the subsequent disease risk [6]. A further exploration of the interaction between dietary fats, obesity, and metabolic inflammation within the context of gastrointestinal cancers is necessary. A better understanding is urgently required to understand how different dietary components may regulate metabolic inflammation within the context of obesity-driven cancers. This will allow for a more complete understanding of the potential role that obese adipose tissue and precision nutrition approaches have on metabolic inflammation in gastrointestinal cancer oncogenesis.

2. Obesity-Related Metabolic Triggers and Gastrointestinal Cancer

Obesity-associated metabolic dysfunction and chronic low-grade inflammation predispose people to metabolic disease development, including T2D, non-alcoholic fatty liver disease (NAFLD), and CVD [7]. Excess body fat elicits several metabolic characteristics, including hyperinsulinaemia, IR, hyperglycaemia, hypercholesterolaemia, elevated non-esterified fatty acid (NEFA, or free fatty acid) levels, and elevated triacylglycerol (TAG) levels [8], which may have oncogenic implications. These metabolic abnormalities can drive tumorigenesis through dysregulation in multiple signalling pathways (Figure 1). In obesity, plasma insulin increases with glucose levels due to the heightened insulin secretion, paired with decreased insulin clearance [9,10]. Insulin is oncogenic through activation of the phosphoinositide-3-kinase (PI3K)/Akt signalling pathway, which increases carcinogenesis in breast and colon cancer cells [11]. The PI3K/Akt signalling pathway acts by the PI3K enzyme activating Akt which subsequently activates target proteins, the main one being the serine/threonine kinase mechanistic target of rapamycin (mTOR), to promote cellular growth, proliferation, and invasion [12]. The PI3K/Akt downstream effectors concerning oncogenesis have previously been reviewed [12]. mTOR regulates cell growth through the phosphorylation of targets which control protein anabolism, growth factor signalling, and nutrient metabolism [13]. An increase in insulin receptor expression is a poor prognostic factor in lung, breast, and colon cancer [14,15,16]. Activation of the insulin receptor initiates the downstream activation of PI3K/Akt, mTOR, and rat sarcoma (RAS)-mitogen-activated protein kinase (MAPK) pathways, all of which are associated with cell survival and proliferation [17]. The RAS–MAPK pathway is common in human cancer through the aberrant activation of receptor tyrosine kinase or through gain-of-function mutations primarily seen in the RAS gene [18]. The activation of the MAPK cascade increases cell proliferation, differentiation, and motility through diverse mechanisms, and these were reviewed extensively [18]. Insulin-like growth factor 1 (IGF-1) signalling is also implicated in cancer development. IGF-1 increased the proliferation in oesophageal adenocarcinoma (OAC) cells and was higher in the serum of viscerally obese OAC patients [19]. Additionally, IGF-1 levels are higher in the serum of colorectal cancer (CRC) patients [20]. IR is characterised by insulin-dependent tissues being unable to take up and utilise glucose efficiently via glucose transporter type 4 (GLUT4) [21], resulting in hyperglycaemia. Warburg first observed that increased blood glucose was associated with tumorigenesis [22]. The Warburg effect occurs in proliferating cells and tumours where the glucose uptake rate increases, paired with lactate generation even when there is normal mitochondrial function and ample oxygen availability [23]. This is thought to allow for quick adenosine triphosphate (ATP) synthesis, an increase in biosynthetic pathways and cell signalling, and the disruption of tissue architecture, all enhancing tumorigenesis. The excess blood glucose seen in obesity supports the increased energetic demand of cancer cells. This is achieved through multiple mechanisms including increased insulin/IGF-1, pro-inflammatory cytokines, and pro-survival Akt/mTOR signalling [24].
Cancer cells require cholesterol for membrane synthesis and cholesterol metabolites are required for cell proliferation, migration, and invasion. Excess cholesterol increases intestinal stem cell proliferation to promote tumorigenesis [25]. Additionally, obese adipose tissue also secretes other pro-tumorigenic hormones including estrogen, leptin, reactive oxygen species (ROS), and cytokines [26]. Ultimately, obesity leads to a plethora of metabolic abnormalities with oncogenic capabilities, which could drive tumorigenesis through multiple cellular signalling pathways (Figure 1).

2.1. Adipose Tissue Heterogeneity and Inflammation

Heterogeneous obesity phenotypes are caused by differences in adipose location, structure, variable adipocyte, and immune cell infiltration and function [27]. There are two main adipose depots, subcutaneous adipose tissue (SAT) and visceral adipose tissue (VAT). It is proposed that VAT is the main source of obesity-driven inflammation, releases more fatty acids, and develops higher IR in comparison to SAT, providing a greater risk for developing metabolic dysfunction [28,29]. VAT’s proximity to the gastrointestinal organs, specifically the stomach, liver, oesophagus, and colon, may make it particularly problematic. Preclinical and human studies show that the depot origin dictates the lipid metabolism, with VAT displaying higher lipolysis and lipogenesis in comparison to SAT [30,31,32]. Additionally, subpopulations of adipocytes show heterogeneity through their response to external stimuli, including tumour necrosis factor alpha (TNFα), insulin, and human growth hormone [33]. Sexual dimorphism and genetic variance play a role in body fat distribution. Women store adipose tissue predominantly subcutaneously versus viscerally, while men are the opposite [34]. Transcriptomic studies show VAT and SAT adipocytes have many genetic differences in developmental genes and other resident cell populations [35,36]. However, whether gene differences are the cause or consequence of fat distribution patterns is unclear. VAT is known to be more deleterious than SAT in metabolic disease [37]; however, their roles in cancer are less evident.

2.1.1. SFA and MUFA in Adipose Tissue Distribution

Diet may play an important role in adipose distribution (Figure 2). Within the human diet, the predominant SFA is palmitate, while the major MUFA is oleate. Interestingly, pre-adipocytes from human VAT and SAT displayed differential lipid accumulation following treatment with palmitate versus oleate. Acute feeding studies showed that feeding palmitate increases the lipid accumulation in VAT to a greater extent than in SAT in young men, while oleate increases lipid accumulation in SAT and not in VAT [38]. In humans, the replacement of a SFA-rich diet with MUFA showed a decrease in body and fat mass without a decrease in total energy or fat intake [39]. Furthermore, a MUFA-rich diet reduced the visceral adiposity compared to other fatty acids [40]. Ultimately, SFA-rich diets drive a more visceral adiposity, which is linked to more metabolic disease and possible tumorigenic opportunities.

2.1.2. Adipose Tissue Morphology

Adipose morphology or adipocyte architecture also affects the functionality of adipose tissue. Thus, the mechanism through which adipocytes and adipose tissue expand can also dictate metabolic health, and therefore also cancer (Figure 3). Adipogenesis is the process through which adipocytes develop from stem cells and accumulate in adipose tissue. Adipose tissue can expand through an increase in the existing adipocyte size (hypertrophy), or new adipocyte formation (hyperplasia). Hypertrophic adipose tissue is more metabolically unhealthy with an increase in insulin resistance and inflammation independent of BMI [42,43,44]. Alternatively, hyperplastic adipose tissue can be characterised as metabolically healthy, containing smaller adipocytes and reduced blood vessels. The adipogenic potential is disrupted within obese Individuals; this is ascribed in part to the inflammatory cytokine milieu, resulting in lower adipogenic gene expression leading to the formation of larger adipocytes, which are associated with IR, inflammation, and redox stress [45,46,47].

2.1.3. The Impact of SFA and MUFA on Adipose Morphology

Fatty acid composition also has differential effects on adipose expansion. Pre-clinical studies show that feeding a SFA high-fat diet results in a hypertrophic adipose profile, compared to feeding a MUFA high-fat diet, despite an equal weight gain and adipose tissue weight [41] (Figure 2). This observation was mediated via the differential interleukin (IL)-1β-induced expression of peroxisome proliferator-activated receptor gamma (PPARγ) and peroxisome proliferator-activated receptor gamma coactivator 1-alpha (PGC-1α), resulting in adipose tissue hyperplasia following the MUFA diet or hypertrophic adipose after feeding on the SFA diet [41]. Ex vivo fatty acid treatments showed that a specific MUFA, palmitoleate, increases the adipose progenitor proliferation through increasing IGF-1 sensitivity [48]. Other pre-clinical studies show that while hypertrophy is strongly correlated with diet, hyperplasia adipose may be more dependent on the interaction between diet and genetics. The extent of hyperplastic expansion was dependent on genetic strain undergoing a high-fat diet [49]. Overall, adipose expansion mechanisms which cause hypertrophy or hyperplasia may be fatty acid-dependent, thus dictating the metabolic phenotype and potential tumorigenic microenvironment.

2.2. Obesity’s Role in Immune Cell Fractions and Function in Cancer

Adipose tissue was originally viewed solely for energy storage; however, it is now proven to have important endocrine functions including secreting cytokines and adipokines [9]. Adipose tissue is mainly composed of fat-storing adipocytes, which are enveloped by a stromal vascular fraction (SVF) composed of a diverse collection of cells including pre-adipocytes, fibroblasts, endothelial, and immune cells. The expansion of adipose tissue depots enhances immune cell infiltration and instigates more pro-inflammatory immune cell populations to further increase inflammation [10,11]. Adipose immune cell infiltration is probably instigated by hypertrophic adipocytes producing monocyte chemotactic protein-1 (MCP-1), which recruits pro-inflammatory cells [50,51]. With increasing adiposity, the SVF becomes enriched with macrophages, T cells, B cells, dendritic cells (DCs), invariant natural killer T cells (iNKT), mucosal-associated invariant T (MAIT) cells, gamma delta (γδ) T cells, and innate lymphoid cells [52]. The presence of these immune cells is integral to chronic low-grade inflammation development which is essential to metabolic disease and is believed to play an important role in obesity-related cancer risk [12,13].
With increasing adiposity, the presence of adipose tissue macrophages (ATMs) increases, paired with a phenotypic switch from anti-inflammatory M2-like (F4/80-) ATMs to pro-inflammatory M1-like ATMs (F4/80+) [50]. M1-macrophages are central cells in promoting inflammation in the adipose tissue microenvironment, specifically in obesity [53,54,55]. Furthermore, increased NEFA can enhance macrophage polarisation towards the M1-like phenotype [56]. Obese ATMs support tumorigenesis through IL-6 secretion, which promotes stem-like properties. Furthermore, weight reduction in a pre-clinical model reverses the macrophage reprogramming and oncogenesis [57]. ATMs from obese patients induce inflammation and lipid accumulation in cancer cells. Furthermore, tumour-associated macrophages have gene expression profiles more similar to obese ATMs versus lean ones [58].
T cell populations, specifically CD4+ and CD8+, are also changed in obesity. In the adipose, CD8+ T cells experience higher activation, while T cell subsets shift to a more pro-inflammatory phenotype with higher T helper (Th) subsets of Th1 and Th17 cells and lower levels of regulatory T (Treg) and Th2 cells [59]. Alternatively, tumour resident CD8+ T cells were reduced in tumours from obese mice. Additionally, tumour infiltrating CD8+ T cells were functionally and metabolically impaired with lower chemokine secretion and proliferation capability, resulting in a reduced ability to control tumour growth [60]. Metabolic plasticity has been identified as a pivotal regulator of T cell responses, with Treg cells exhibiting heightened employment of fatty acid oxidation (FAO), whereas effector cells preferably use glycolysis [61]. Remarkably, adipose tissue procured from viscerally obese patients have increased secreted levels of mediators related to Th17 immune responses. Th17 and Treg cell populations are balanced in the gastrointestinal system, resulting in normal immune system function and tissue homeostasis. At the gastric tumour site, Th17 and Treg cells infiltrate, proving that the tumour microenvironment could cause the Th17 and Treg cells to become imbalanced [62].
iNKT cells are a particular subgroup of T cells which are swiftly activated in response to excess lipids bound through CD1d, an antigen-presenting molecule which is expressed by DCs or macrophages [63]. Interestingly, CD1d cells are highly expressed in the omentum, an integral part of the VAT. However, the frequencies of these cells are depleted in the omental VAT of morbidly obese patients and cancer patients [64]. Due to the iNKT cell’s close interplay with lipid antigens, it is foreseeable that lipid profile modifications in the lipid profile of the tumour microenvironment would alter their immuno-modulatory effects. Increased lactic acid levels (which are indicative of glycolytic metabolism) in the tumour microenvironment have been implicated in reducing PPARγ on intratumoural iNKT, diminishing cholesterol synthesis and IFN-γ production and reducing their anti-tumour immunity efficacy [54]. However, the introduction of a PPARγ agonist combatted these effects and restored interferon gamma (IFN-γ) production [65]. This indicates the significance of the tumour microenvironment’s lipid profile to promote an effective anti-tumour immune response.
Following high-fat diet initiation, B cells infiltrate into the adipose tissue [66]. B cells secrete pro-inflammatory IL-6, IL-8, and TNFα while inducing other cells to secrete leptin and MCP-1, which are related to intracellular pathways that promote CRC growth and metastatic spreading [67,68]. Additionally, B cells can modulate T cell behaviour, perpetuating inflammation and insulin resistance [69,70,71]. Within the tumour microenvironment, B cells recruit and activate T cells which influence other immune cells to resist tumour cells [72]. Regulatory B cells (Bregs) produce anti-inflammatory IL-10, IL-35, and transforming growth factor—beta (TGF-β), which inhibits immunity resulting in promoting tumour growth [73]. Additionally, Bregs can deplete CD8+ T cells, further increasing the immunosuppression [74].
DCs play a fundamental role in antigen presentation and commencing the anti-tumour immune response, and have been identified as a prominent player in obesity-associated immune responses. DCs represent a significant proportion of infiltration cells during adipose expansion [75]. Additionally, high NEFA levels lead to lipid-loaded DCs with diminished antigen-presenting capabilities and a decreased capacity to effectively stimulate T cells [76]. Normal DC function is required for T cell-mediated tumour clearance. DC-dependent immunotherapy reduced the tumour size in lean mice but was greatly reduced in obese mice [77].
Myeloid-derived suppressor cells (MDSC) increase with obesity in mouse models and humans, in circulation and within adipose tissue [78,79]. Intriguingly, lipid accumulation at the tumour site has been related to metabolic plasticity in MDSCs, guiding them from a glycolytic phenotype towards the enhanced utilisation of FAO and oxidative phosphorylation. This metabolic preference shift confers MDSCs with enhanced immunosuppressive properties, leading to a diminished effect on anti-tumour immunity [80]. High-fat diet-enhanced MDSC accumulation results in increased tumour progression and metastasis through reduced T cell activation [81].
Neutrophils have been reported to be increased in morbidly obese humans [82]. Interestingly, within a glucose-limited tumour microenvironment, neutrophils use FAO to fuel ROS production and suppress T cells [83]. This metabolic plasticity has been implicated in aiding cancer growth, metastasis, and recurrence [84]. Additionally, a high-fat diet elevates levels of granulocyte-macrophage colony-stimulating factor (GM-CSF), which increases neutrophil presence and promotes tumour growth and metastasis [85].
Obese humans have diminished natural killer (NK) cell frequencies with decreased cytotoxicity, which may lead to impaired tumour immune responses [86]. A fatty acid-enriched microenvironment impairs NK cell functionality [87], validating the theory that obesity may have disadvantageous effects on NK cell performance. NK cells from obese cancer patients are recruited to adipose tissue where they undergo irreversible dysregulation leading to cell death. In OAC patients, higher NK cell frequencies have been reported within the VAT whilst diminished expression was detected within tumour tissue [88,89].

2.3. Diet and Fatty Acid’s Role in Immune Cell Fractions

Obesity-associated inflammation and metabolic perturbations are partly caused by the alterations in adipose immune cell phenotypes alluded to above and previously reviewed [52,90]. Recent evidence suggests that the extent of these changes can be dependent on the composition of fatty acids that cell populations are exposed to (Figure 2). Fatty acids can be sourced from the diet, as well as resulting from endogenous de novo lipogenesis in response to energy excess and fatty acid metabolism. The pro-inflammatory effects of SFA are well-characterised. Briefly, SFA can signal through a cytosolic lipid-responsive pattern recognition receptors protein complex, the NOD-, LRR-, and pyrin domain-containing protein 3 (NLRP3) inflammasome, promoting pro-inflammatory cytokine expression including IL-1β and IL-18. This occurs through a two-phase process requiring stress signals which can include cholesterol, NEFA, ATP, pathogens (e.g., lipopolysaccharide (LPS)), glucose, and ROS, all of which are increased in obesity [6]. The SFA palmitate activated the NLRP3 inflammasome in both macrophages and DCs in mice following a high-fat diet [41,91,92,93,94]. Additionally, a palmitate treatment increased TNFα and IL-1β secretion, paired with a decrease in anti-inflammatory cytokine IL-10 secretion [93,95]. Conversely, MUFA does not activate the NLPR3 inflammasome like SFA. MUFA reverses pro-inflammatory cytokine expression following a SFA diet. This can be attributed to a higher level of anti-inflammatory gene expression (IL-10, macrophage galactose N-acetyl-galactosamine specific lectin 2 (Mgl2), mannose receptor C-type 1 (Mrc1), Tgfb1) and a shifted macrophage phenotype towards M2, as displayed by an increase in the oxygen consumption rate [96].
In macrophages, palmitate has been reported to elevate pro-inflammatory response signals [97,98]. Palmitate increases the expression of NLRP3, nitric oxide (NO), IL-1β, IL-6, TNFα, nuclear factor kappa B (NF-κB), c-Jun N-terminal kinases (JNK), mitogen-activated protein kinase kinase (MKK)4/7, IL-10, MCP-1, IFN-γ, M1 polarisation, and CD36 [98,99,100]. In contrast, oleate has anti-inflammatory effects by inhibiting the pro-inflammatory responses driven by SFA steric acid [101], along with promoting M2-like polarisation in macrophages [102]. In T cells, it has been demonstrated that palmitate increases the expression of insulin receptors, ROS, and cytokines (IFN-γ, IL-1β, IL-2, IL-6, IL-8, IL-10, TNFα), and insulin receptor substrate 1 (IRS-1) generation and proliferation [103]. Additionally, palmitate increases the expression of PI3K/Akt, JNK, and extracellular signal-regulated kinase (ERK)1/2 pathways [104]. Alternatively, oleate decreases proliferation and pro-inflammatory cytokines (IL-2, IFN-γ) [105]. In B cells, palmitate reprogrammed B cells to be immunosenescent [106]. Oleate was required for normal mTOR activity and mitochondrial function, and to prevent endoplasmic reticulum (ER) stress [107]. Palmitate stimulates the pro-inflammatory secretion of IL-1β through toll-like receptor (TLR) activation in DCs [92,108]. In neutrophils, palmitate increases ERK1/2, Akt, ROS, and chemotaxis [109,110]. Oleate decreases migration while also increasing ERK1/2, Akt, NF-κB, ROS, IL-1β, IL-8, and ATP [111,112,113,114,115]. The replacement of palmitate with oleate can reduce pro-inflammatory cytokine secretion. Since palmitate and oleate have different effects on adipose location, expansion, and immune cell characteristics, they could also have differential effects on obesity-related cancer initiation and progression.

3. Metabolic Flexibility in Cancer Cells in the Tumour Microenvironment

Extensive metabolic reprogramming occurs in cancer to keep up with the increased energy demand and to obtain membrane materials for proliferation. Metabolic reprogramming of fatty acids is activated in cancer cells to support their increased bioenergetic demand [116]. These changes in fatty acid metabolism are often paired with mitochondrial dysfunction, which is common in cancer cells to aid tumour progression [117]. Aberrant cellular metabolism is an essential survival advantage induced by cancer to escape the cytotoxic effects of chemotherapy and chemoradiotherapy [118,119]. Lipid metabolism, particularly FAO, enhances treatment resistance in cancer cells through the upregulation of lipogenic or lipolytic enzyme expression [120]. Resistant cells often increase sterol regulatory element binding protein (SREBP)-induced de novo lipogenesis through fatty acid synthase (FASN), and the elongation of very long-chain fatty acid 6 (ELOVL6) or stearoyl-CoA desaturase 1 (SCD1) overexpression in tyrosine kinase inhibitor-resistant cells [121,122]. Additionally, mitogen-activated protein kinase pathway inhibitors increase FAO, which can result in drug resistance [123]. The exploitation of lipid metabolism and FAO is strongly elicited by metastatic cancer cells, M2 macrophages, memory CD8+ T cells, and tissue-resident Treg cells, specifically in VAT [124,125,126,127].
Another metabolic mechanism that cancer cells utilise is glucose metabolism reprogramming, as an innate adjustment in cancer cells. Glucose normally undergoes aerobic respiration, resulting in pyruvate production, which is further converted into acetyl-CoA. The conversion of acetyl-CoA to malonyl-CoA then enables the endogenous production of the SFA palmitate within the cell. From here, palmitate can be desaturated by SCD1 and converted into palmitoleate, or can be elongated by ELOVL6 into stearic acid and then desaturated by SCD1 to form the MUFA oleate. These fatty acids are stored in lipid droplets to be utilised for energy demands during times of cellular stress [128]. Remarkably, FASN is diminished in obesity [129,130], whilst FAO is elevated [131], indicating that obesity may encourage a metabolic shift resulting in the enhanced utilisation of lipid metabolism and FAO. Additionally, FASN is thought to modulate thresholds that trigger receptor signalling and ultimately regulate the balance between anti-proliferation and tumorigenesis [132]. FAO can also be activated by other upstream activators such as AMP kinase (AMPK), promyelocytic leukaemia-peroxisome proliferator-activated receptor (PML-PPAR) pathway, and glycolysis [133]. These cellular metabolic aberrations which increase cancer cell formation can be extended to lead to a metastatic phenotype.
Primary cancer cells rely heavily on glycolytic metabolism to grow and survive, prompting this metabolic dependence despite the oxidative stress prompted by radiation-induced inflammation to facilitate DNA damage repair [134]. However, metastatic cancer cells depend more on oxidative phosphorylation-associated metabolism and FAO [135,136]. Energy metabolism flexibility is critical in aiding cancer cells’ ability to undergo epithelial–mesenchymal transition (EMT) and migration to facilitate distant metastasis [137].

4. Obesity and Cancer Metastasis

Research is centred on the impact of obesity and its contribution to the metabolic changes and EMT that support cancer cells’ migratory capacity to develop distant metastasis. Metastasis encompasses a succession of transfiguring alterations in cancer cells including their metabolic preferences and plasticity, and in their surrounding stroma, which can be triggered as a response to cancer therapies. The phases entailed in the colonisation of distant metastasis include local invasion, intravasation, circulation of cancer cells, extravasation, and, lastly, the establishment of local and distant metastasis [138]. Obesity-linked cancers, including gastric and colon cancer, often metastasise to the omentum, which advocates the contribution of adipose tissue to the metastatic cascade [139]. Adipose tissue can drive this cascade through multiple mechanisms which include increased adipokine/cytokine secretion, metabolic reprogramming, and angiogenesis. An increased secretion of IL-6, leptin, IGF-1, and TNFα in obesity promotes EMT and inflammation, and dampens the immune response [2]. Furthermore, angiogenesis is promoted in obese adipose through an increase in the release of pro-angiogenetic TNFα, IL-6, and vascular endothelial growth factor (VEGF) [140]. These advancements position obese adipose tissue as a highly favourable site to facilitate the development of pre-metastatic niches [141]. Ultimately, adipose tissue can increase tumorigenesis and aid the metabolic flexibility required by cancer cells and metastasis.

5. Evidence Linking Obesity to Gastrointestinal Cancer Risk

Adipose tissue expansion, specifically VAT, can affect gastrointestinal tract cancers due to anatomical proximity. Upper gastrointestinal cancers such as oesophageal tumours are exposed to acid reflux and bile acid following excess dietary fat ingestion [142]. Lower gastrointestinal cancers, including CRC, are surrounded by VAT, increasing the exposure to a sub-acute pro-inflammatory environment [143]. The dysregulation of immune cells in the adipose tissue, discussed above, may also create an environment that drives tumorigenesis. Increasing adipose inflammation with obesity disrupts tissue homeostasis, hampers immunological responses, and can lead to tissue hyperplasia or death, ultimately creating obesity-driven tumorigenesis [143,144,145,146]. Cytokines that are classically upregulated in obesity, such as TNFα, IL-6, and TGFβ, promote tumour cell proliferation and invasion, and possible tumour formation [147,148]. Furthermore, the phenotypic changes described above in CD4+ and CD8+ cells are involved with tumour growth and metastasis proximal to adipose tissue [149]. Obese adipose tissue immune cell dysfunction creates an environment which allows tumorigenic cell growth and metastasis. The culmination of this results in obesity instigating cancer including OAC, gastric, liver, and CRC (Figure 4).

5.1. Obesity and OAC

Large-scale epidemiological studies consistently illustrate a compelling association between the risk of cancer onset or progression and increased BMI for numerous gastrointestinal cancers including OAC. OAC is among one of the cancer types most strongly correlated with escalating obesity levels [150,151,152,153,154], making it an exemplary model for studying obesity’s influence on cancer, especially because of its proximity to VAT depots. Interestingly, oesophageal cancer adipose tissue has been reported to recruit immune cells while negatively impacting their function, thereby enhancing anti-tumour immunity [88,155,156,157]. Previous research has reported that various pro-inflammatory mediators in the circulation and expressed within the tissue have shown associations with clinical outcomes in OAC, particularly factors that are involved in the recruitment and activation of innate immune cells [158]. Adipose tissue energy metabolism and the impact of its secretome on cancer cell metabolism is an emerging area of research. In OAC patients, VAT had higher oxidative phosphorylation compared to SAT. Additionally, VAT secretions increase angiogenic and inflammatory cytokines including VEGF-A, VEGF-C, IL-2, IL-16, and TNFα [159]. Viscerally obese OAC patients with increased oxidative phosphorylation were correlated with metabolic dysfunction and increased pro-inflammatory mediators IL-5 and IL-7. Furthermore, glutamine levels are reduced while its metabolised product glutamate’s levels are increased in the adipose secretome of obese compared to non-obese OAC patients [160]. Recent research has indicated that the secretome of adipose explants derived from OAC patients is altered due to increased visceral adiposity [157]. Interestingly, inflammatory factors including Eotaxin-3, MCP-1, macrophage-derived chemokine (MDC), and IL-17 were shown to be increased in the adipose secretome of patients with enlarged VAT depots. Previously, these factors have been shown to increase immune cell infiltration [161] and may be linked with maintaining the low-grade inflammatory state that is associated with obesity.

5.2. Obesity and Gastric Cancer

Overweight and obesity cause approximately 6% of gastric cancer (GC) [162]. Furthermore, GC metastasis is commonly directed towards the VAT depot, highlighting the important role adipose plays in tumour progression. The human gastric adenocarcinoma cell proliferation and migration rate increased following incubation with human visceral adipose-conditioned media (ACM). Additionally, the S phase population of the cell cycle was increased. Furthermore, GC cells cultured with human visceral ACM were injected into nude mice, which increased the rate of tumour growth compared to cells not grown in ACM [163]. In vitro, the co-culture of adipocytes with GC cells drove the adipocytes to dedifferentiate into cancer-associated fibroblasts, with increased IL-6 secretion. Furthermore, VAT proximal to primary tumours displayed reduced adiponectin levels in patients who exhibited subserosal or serosal invasion [164]. Visceral ACM induced the angiogenesis of GC cells through Akt phosphorylation and overexpression of VEGF-A with increased secretion of chemokine (C-X-C motif) ligand (CXCL) 2 [165]. The increased expression of fatty acid transporters such as CD36 and fatty acid-binding protein 1 (FABP1) are increased in obesity [166,167] and drive increased pathway expression, which increases GC metastasis. This highlights the effects of obese adipose on the tumorigenic environment in the stomach.

5.3. Obesity and Liver Cancer

Liver cancer, or hepatocellular carcinoma (HCC), has a constantly increasing trend in the USA and many European countries. HCC can arise from liver cirrhosis, credited to hepatitis B and C virus infections and/or heavy alcohol intake [168]. The increase in HCC is paralleled with an increase in NAFLD. One study found that NAFLD, in the absence of obesity, elevated the cancer risk primarily in the liver, gastrointestinal tract, and uterus [169]. Conversely, approximately 23% of HCC cases in the UK are caused by overweight and obesity [162]. In male HCC patients, the VAT depot mass was higher in HCC versus non-HCC patients and it was a risk factor for the recurrence of HCC after liver transplantation [170]. Adipose exclusively secretes adiponectin, an adipokine which reduces triglycerides levels and controls insulin signalling, whose levels are decreased in obesity [171]. Interestingly, adiponectin secretion is positively correlated with a poor prognosis of liver cancer [172]. This may indicate that obesity alone is not the only determinant factor in HCC prognosis. Inflammatory adipose-derived cytokines, including TNFα and IL-6, are oncogenic signalling molecules in liver cancer [173]. Pro-inflammatory activities induced by the adipokine leptin and lipotoxicity, reflecting increased fatty acid storage that spills over or escapes from adipose, increase the proliferation and oncogenic mutations resulting in carcinogenesis in the liver [174]. The second most prevalent liver tumour is cholangiocarcinoma, which occurs within the bile duct epithelium [175]. A high BMI is significantly associated with increased tumour size and metastasis rates leading to a poor prognosis and a heightened risk of recurrence. Furthermore, the tumour tissue from obese patients displayed altered immune characteristics which included increased PD-L1 expression, decreased CD8+ T cells, and increased FOX p3 + T cells [176]. Ultimately, the obese adipose phenotype has significant effects on liver health and can perpetuate hepatic cancer formation.

5.4. Obesity and CRC

CRC is a predominantly obesity-associated cancer which is strongly associated with lifestyle factors such as diet [177]. Obese-associated inflammation can promote CRC. Mutagenesis can occur through increased reactive oxidative damage and epigenetic silencing [142,178]. Increased IL-6 production can shift intestinal macrophages towards a M2-like macrophage phenotype in mouse models which overlaps with tumour-associated macrophages [179]. These M2-like macrophages then recruit B cells and γδ T cells to the tumour environment. CC motif chemokine receptor 6 (CCr6)-expression γδ T cells secrete IL-17, which further increased the colon inflammation in a T cell receptor alpha (TCRα) −/− mouse model [180]. When the recruitment of B cells and γδ T cells was blocked, there was a suppression of CRC development [181]. ACM from obese patients and lean and obese CRC subjects released more IL-6, IL-8, and MCP-1 compared to healthy lean subjects. This ACM was then cultured on DCs, promoting the differentiation and increased expression of programmed death-ligand 1 and 2 (PD-L1, PD-L2) with a diminished IL-12/IL-10 ratio, thus preventing DC-mediated γδ T cell activation [182]. While obesity may drive CRC development, it can also have differing effects on disease outcomes. There is a negative impact of BMI concerning disease relapse and death in stage III patients with a BMI > 30 km m2 [183]. Alternatively, there was also the emergence of the obesity paradox. Immune checkpoint therapy has been reported to have a positive association with obesity [184]. This highlights the ambiguous nature of obesity-driven cancers and therapies, which requires a greater understanding of disease mechanisms leading towards a personalised response.

6. SFA and MUFA’s Roles in Gastrointestinal Tumorigenesis

Lipids are emerging as key molecules fuelling cancer cell proliferation. The nutritional modulation of dietary fat is now thought to be important in cancer; however, little is known about how individual dietary lipids may regulate tumour growth and metastasis. The overconsumption of dietary fat may be positively or negatively correlated with cancer risk, depending on the fatty acid and cancer type. In vitro, the SFA lauric acid suppressed CRC cell proliferation, while in vivo, a palmitate-rich high-fat diet stimulated tumour growth. In vitro, MUFA oleate promoted growth in colon cancer cell lines while suppressing the growth and survival of GC cells [185]. Research into the role of dietary SFA and MUFA in tumorigenesis is increasing. Cell proliferation pathways, including ERK1/2-mTOR-NF-κB and PI3K/Akt, were differentially modulated by SFA and MUFA in a model- and cancer-dependent manner [185], suggesting that different dietary lipids may have distinct effects in tumorigenesis.
There is some evidence, albeit limited, that fatty acids may have distinct effects on gastrointestinal cancers. Palmitate upregulated carnitine palmitoyltransferase 1A (CPT1A) in the disease sequence from Barrett’s oesophagus to OAC, in both in vitro and mouse models, resulting in increased cell proliferation [186]. Importantly, CPT1A is a rate-limiting enzyme in FAO, whose substrate is palmitate, which has been linked with promoting cancer cell proliferation [187]. Oleate downregulated cell proliferation in OE19 and OE33 oesophageal cancer cell lines through the increased phosphorylation of AMPK with reduced S6 activation. Additionally, oleate increased the expression of tumour suppression genes p53, p21, and p27 [188]. Palmitate promoted metastasis both in vitro and in vivo through CD36 receptor activity and via the AKT/glycogen synthase kinase 3 beta (GSK3β)/β-catenin pathway [24]. A palmitate treatment also promoted gastric metastasis through the fatty-acid binding protein 5-specific protein 1-urothelial cancer-associated 1 (FABP5/SP1/UCA1) pathway [189]. In vitro, a co-culture of GC cells with isolated omental adipocytes showed an increase in oleate within the gastric cell. An oleate treatment on gastric cells enhanced the invasiveness through the PI3K/Akt pathway [190]. In human hepatoma cells, palmitate disturbed lipid metabolism and increased the protein expression of NLRP3 inflammasome and ER stress, while oleate was able to rescue these cells from pyroptosis. In vivo regression studies showed that the replacement of a high-fat diet with an oleate-rich olive oil reduced liver abnormalities and inhibited ER stress [191]. In vitro, the administration of palmitate to CRC cells increased proliferation through an enhanced expression of β2-adrenergic receptors, which are vital for CRC growth [192]. An in vitro palmitate treatment of intestinal organoids increased the number of leucine-rich repeating-containing receptor 5 (Lgr5+) intestinal stem cells in a PPAR-d dependent manner, which boosted their ability to form colorectal adenocarcinomas [193]. High-fat diet feeding in CRC mice showed an increase in palmitate, which increased the beta-2 adrenergic receptor (β2AR) expression and β-adrenergic signalling pathway, which was then reduced with the removal of the high-fat diet [192]. Furthermore, a short-term palmitate-rich diet induced a more aggressive tumour cell profile that endured as cellular memory in a CD36-dependent manner [194]. The blockage of CD36 expression inhibited metastasis, highlighting that dietary fatty acids are needed to promote metastasis (Figure 4). Whilst most of the data discussed above pertain to in vitro fatty acid exposures and pre-clinical in vivo diet-induced mechanisms, we need greater translational data to understand if, ultimately, diets rich in saturated fat, such as palmitate, may contribute to cancer development to a greater extent than unsaturated fat, including oleate.

7. Conclusions

There are clear associations between obesity and gastrointestinal cancers. However, obesity is extremely heterogeneous and highly dictated by adipose structure, immune cell fractions, and dietary components. Hypertrophic adipose mass recruits more pro-inflammatory immune cells compared to hyperplastic adipose mass. Furthermore, visceral adipose has more pro-inflammatory behaviour and fatty acids released compared to subcutaneous adipose, which exacerbates cellular metabolic dysfunction. The importance of dietary fats, specifically SFA and MUFA, on the adipose function and extent of inflammation is potentially evident. SFA increases hypertrophic inflammatory adipose microenvironments, creating ideal pre-metastatic niches for tumorigenesis to occur. Alternatively, MUFA does not display these same effects. Furthermore, the different effects palmitate and oleate have on tumorigenesis in gastrointestinal cancers highlight the importance of the specific type of fatty acid intake on cancer initiation and/or progression. However, further research on the difference between SFA and MUFA in gastrointestinal cancers is required to fully elucidate the mechanisms that differ between these fatty acids. This will give us greater translation insight with respect to the true impact of dietary fat intake and risk. With this enhanced knowledge base, hopefully we can develop dietary preventions and/or interventions. These will rely on a more targeted understanding of obesity and the nutrition environment, embracing a precision nutrition approach which may be a more effective line of treatment. Both obesity and cancer are highly diverse and individualised diseases on their own, and their complexity may be amplified when they are combined. Precision nutrition therapy could target individual conditions to maximise the effectiveness and is a promising tool in cancer therapy.

Author Contributions

K.A.J.M. and F.O. wrote the manuscript. J.O. and H.M.R. read, revised, and approved the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

K.A.J.M., J.O. and H.M.R. were funded by Precision Oncology Ireland (18/SPP/3522), a Science Foundation Ireland Strategic Partnership Programme. H.M.R. was funded by the Science Foundation Ireland Frontiers Award Programme (SFI 19/FFP/6625). F.O. and J.O. were funded by Breakthrough Cancer Research (209740/16104).

Acknowledgments

We would like to thank Precision Oncology Ireland, Science Foundation Ireland, and Breakthrough Cancer Research for supporting our work. The authors used BioRender to generate the figures.

Conflicts of Interest

The authors declare no conflict of interest.

Glossary

ACMadipose-conditioned media
AMPKAMP-activated protein kinase
ATMadipose tissue macrophage
ATPadenosine triphosphate
β2ARbeta-2 adrenergic receptor
BMIbody mass index
Bregregulatory B cells
CCr6CC motif chemokine receptor 6
CPT1Acarnitine palmitoyltransferase 1A
CRCcolorectal cancer
CVDcardiovascular disease
CXCLchemokine (C-X-C motif) ligand
DCdendritic cell
ELOVLelongation of very-long-chain fatty acid
EMTepithelial–mesenchymal transition
ERendoplasmic reticulum
ERKextracellular signal-regulated kinase
FABP1fatty-acid binding protein 1
FABP5/SP1/UCA1fatty-acid binding protein 5-specific protein 1-urothelial cancer associated 1
FAOfatty acid oxidation
FASNfatty acid synthase
γδgamma-delta
GCgastric cancer
GLUT4glucose transporter type 4
GM-CSFgranulocyte-macrophage colony-stimulating factor
GSK3βglycogen synthase kinase 3 beta
HCChepatocellular carcinoma
IFN-γinterferon gamma
IGF-1insulin-like growth factor 1
ILinterleukin
iNKTinvariant natural killer T cells
IRinsulin resistance
IRS-1insulin receptor substrate 1
JNKc-Jun N-terminal kinases
Lgr5+leucine-rich repeating-containing receptor 5
LPSlipopolysaccharide
MAITmucosal-associated invariant T cells
MAPKmitogen-activated protein kinase
MCP-1monocyte chemoattractant protein-1
MDCmacrophage-derived chemokine
MDSCmyeloid-derived suppressor cells
Mgl2macrophage galactose N-acetyl-galactosamine specific lectin 2
MKKmitogen-activated protein kinase kinase
Mrc1mannose receptor C-type 1
mTORmechanistic target of rapamycin
MUFAmonounsaturated fatty acid
NAFLDnon-alcoholic fatty liver disease
NEFAnon-esterified fatty acid (or free fatty acid)
NF-κBnuclear factor kappa B
NKnatural killer
NLRP3NOD-, LRR- and pyrin domain-containing protein 3
NOnitric oxide
OACoesophageal adenocarcinoma
PD-Lprogrammed death-ligand
PGC-1αperoxisome proliferator-activated receptor gamma coactivator 1-alpha
PI3Kphosphoinositide-3-kinase
PML-PPARpromyelocytic leukaemia-peroxisome proliferator-activated receptor
PPARγperoxisome proliferator-activated receptor gamma
RASrat sarcoma
ROSreactive oxygen species
SATsubcutaneous adipose tissue
scd1stearoyl-CoA desaturase
SFAsaturated fatty acid
SREBPsterol regulatory element binding proteins
SVFstromal vascular fraction
T2Dtype 2 diabetes
TAGtriacylglycerol
TCRαT cell receptor alpha
TGF-βtransforming growth factor beta
Thhelper T cells
TLRtoll-like receptor
TNFαtumour necrosis factor alpha
Tregregulatory T cells
VATvisceral adipose tissue
VEGFvascular endothelial growth factor

References

  1. Engin, A. The Definition and Prevalence of Obesity and Metabolic Syndrome. Adv. Exp. Med. Biol. 2017, 960, 1–17. [Google Scholar] [CrossRef]
  2. Harris, B.H.L.; Macaulay, V.M.; Harris, D.A.; Klenerman, P.; Karpe, F.; Lord, S.R.; Harris, A.L.; Buffa, F.M. Obesity: A perfect storm for carcinogenesis. Cancer Metastasis Rev. 2022, 41, 491–515. [Google Scholar] [CrossRef] [PubMed]
  3. Kyrgiou, M.; Kalliala, I.; Markozannes, G.; Gunter, M.J.; Paraskevaidis, E.; Gabra, H.; Martin-Hirsch, P.; Tsilidis, K.K. Adiposity and cancer at major anatomical sites: Umbrella review of the literature. BMJ 2017, 356, j477. [Google Scholar] [CrossRef] [PubMed]
  4. Bluher, M. Metabolically Healthy Obesity. Endocr. Rev. 2020, 41, bnaa004. [Google Scholar] [CrossRef] [PubMed]
  5. Charles-Messance, H.; Mitchelson, K.A.J.; De Marco Castro, E.; Sheedy, F.J.; Roche, H.M. Regulating metabolic inflammation by nutritional modulation. J. Allergy Clin. Immunol. 2020, 146, 706–720. [Google Scholar] [CrossRef] [PubMed]
  6. Ralston, J.C.; Lyons, C.L.; Kennedy, E.B.; Kirwan, A.M.; Roche, H.M. Fatty Acids and NLRP3 Inflammasome-Mediated Inflammation in Metabolic Tissues. Annu. Rev. Nutr. 2017, 37, 77–102. [Google Scholar] [CrossRef] [PubMed]
  7. Shin, J.A.; Lee, J.H.; Lim, S.Y.; Ha, H.S.; Kwon, H.S.; Park, Y.M.; Lee, W.C.; Kang, M.I.; Yim, H.W.; Yoon, K.H.; et al. Metabolic syndrome as a predictor of type 2 diabetes, and its clinical interpretations and usefulness. J. Diabetes Investig. 2013, 4, 334–343. [Google Scholar] [CrossRef]
  8. Virtue, S.; Vidal-Puig, A. Adipose tissue expandability, lipotoxicity and the Metabolic Syndrome—An allostatic perspective. Biochim. Biophys. Acta 2010, 1801, 338–349. [Google Scholar] [CrossRef]
  9. Kahn, S.E.; Prigeon, R.L.; McCulloch, D.K.; Boyko, E.J.; Bergman, R.N.; Schwartz, M.W.; Neifing, J.L.; Ward, W.K.; Beard, J.C.; Palmer, J.P.; et al. Quantification of the relationship between insulin sensitivity and beta-cell function in human subjects. Evidence for a hyperbolic function. Diabetes 1993, 42, 1663–1672. [Google Scholar] [CrossRef]
  10. Bergman, R.N.; Ader, M.; Huecking, K.; Van Citters, G. Accurate assessment of beta-cell function: The hyperbolic correction. Diabetes 2002, 51 (Suppl. S1), S212–S220. [Google Scholar] [CrossRef]
  11. Tomas, N.M.; Masur, K.; Piecha, J.C.; Niggemann, B.; Zanker, K.S. Akt and phospholipase Cgamma are involved in the regulation of growth and migration of MDA-MB-468 breast cancer and SW480 colon cancer cells when cultured with diabetogenic levels of glucose and insulin. BMC Res. Notes 2012, 5, 214. [Google Scholar] [CrossRef]
  12. Rascio, F.; Spadaccino, F.; Rocchetti, M.T.; Castellano, G.; Stallone, G.; Netti, G.S.; Ranieri, E. The Pathogenic Role of PI3K/AKT Pathway in Cancer Onset and Drug Resistance: An Updated Review. Cancers 2021, 13, 3949. [Google Scholar] [CrossRef]
  13. Laplante, M.; Sabatini, D.M. mTOR signaling in growth control and disease. Cell 2012, 149, 274–293. [Google Scholar] [CrossRef]
  14. Kim, J.S.; Kim, E.S.; Liu, D.; Lee, J.J.; Solis, L.; Behrens, C.; Lippman, S.M.; Hong, W.K.; Wistuba, I.I.; Lee, H.Y. Prognostic impact of insulin receptor expression on survival of patients with nonsmall cell lung cancer. Cancer 2012, 118, 2454–2465. [Google Scholar] [CrossRef] [PubMed]
  15. Gallagher, E.J.; Fei, K.; Feldman, S.M.; Port, E.; Friedman, N.B.; Boolbol, S.K.; Killelea, B.; Pilewskie, M.; Choi, L.; King, T.; et al. Insulin resistance contributes to racial disparities in breast cancer prognosis in US women. Breast Cancer Res. 2020, 22, 40. [Google Scholar] [CrossRef]
  16. Heckl, S.M.; Pellinghaus, M.; Kruger, S.; Bosselmann, C.; Wilhelm, F.; Behrens, H.M.; Schreiber, S.; Rocken, C. Epithelial insulin receptor expression-prognostic relevance in colorectal cancer. Oncotarget 2018, 9, 37497–37508. [Google Scholar] [CrossRef] [PubMed]
  17. Leitner, B.P.; Siebel, S.; Akingbesote, N.D.; Zhang, X.; Perry, R.J. Insulin and cancer: A tangled web. Biochem. J. 2022, 479, 583–607. [Google Scholar] [CrossRef] [PubMed]
  18. Santarpia, L.; Lippman, S.M.; El-Naggar, A.K. Targeting the MAPK-RAS-RAF signaling pathway in cancer therapy. Expert Opin. Ther. Targets 2012, 16, 103–119. [Google Scholar] [CrossRef]
  19. Doyle, S.L.; Donohoe, C.L.; Finn, S.P.; Howard, J.M.; Lithander, F.E.; Reynolds, J.V.; Pidgeon, G.P.; Lysaght, J. IGF-1 and its receptor in esophageal cancer: Association with adenocarcinoma and visceral obesity. Am. J. Gastroenterol. 2012, 107, 196–204. [Google Scholar] [CrossRef]
  20. Ma, J.; Pollak, M.N.; Giovannucci, E.; Chan, J.M.; Tao, Y.; Hennekens, C.H.; Stampfer, M.J. Prospective study of colorectal cancer risk in men and plasma levels of insulin-like growth factor (IGF)-I and IGF-binding protein-3. J. Natl. Cancer Inst. 1999, 91, 620–625. [Google Scholar] [CrossRef]
  21. Perry, R.J.; Samuel, V.T.; Petersen, K.F.; Shulman, G.I. The role of hepatic lipids in hepatic insulin resistance and type 2 diabetes. Nature 2014, 510, 84–91. [Google Scholar] [CrossRef] [PubMed]
  22. Warburg, O. On the origin of cancer cells. Science 1956, 123, 309–314. [Google Scholar] [CrossRef]
  23. Liberti, M.V.; Locasale, J.W. The Warburg Effect: How Does it Benefit Cancer Cells? Trends Biochem. Sci. 2016, 41, 211–218. [Google Scholar] [CrossRef]
  24. Pan, J.; Fan, Z.; Wang, Z.; Dai, Q.; Xiang, Z.; Yuan, F.; Yan, M.; Zhu, Z.; Liu, B.; Li, C. CD36 mediates palmitate acid-induced metastasis of gastric cancer via AKT/GSK-3beta/beta-catenin pathway. J. Exp. Clin. Cancer Res. 2019, 38, 52. [Google Scholar] [CrossRef] [PubMed]
  25. Wang, B.; Rong, X.; Palladino, E.N.D.; Wang, J.; Fogelman, A.M.; Martin, M.G.; Alrefai, W.A.; Ford, D.A.; Tontonoz, P. Phospholipid Remodeling and Cholesterol Availability Regulate Intestinal Stemness and Tumorigenesis. Cell Stem Cell 2018, 22, 206–220.e204. [Google Scholar] [CrossRef]
  26. Nieman, K.M.; Romero, I.L.; Van Houten, B.; Lengyel, E. Adipose tissue and adipocytes support tumorigenesis and metastasis. Biochim. Biophys. Acta 2013, 1831, 1533–1541. [Google Scholar] [CrossRef]
  27. Massier, L.; Jalkanen, J.; Elmastas, M.; Zhong, J.; Wang, T.; Nono Nankam, P.A.; Frendo-Cumbo, S.; Backdahl, J.; Subramanian, N.; Sekine, T.; et al. An integrated single cell and spatial transcriptomic map of human white adipose tissue. Nat. Commun. 2023, 14, 1438. [Google Scholar] [CrossRef]
  28. Wajchenberg, B.L.; Giannella-Neto, D.; da Silva, M.E.; Santos, R.F. Depot-specific hormonal characteristics of subcutaneous and visceral adipose tissue and their relation to the metabolic syndrome. Horm. Metab. Res. 2002, 34, 616–621. [Google Scholar] [CrossRef] [PubMed]
  29. Meza-Perez, S.; Randall, T.D. Immunological Functions of the Omentum. Trends Immunol. 2017, 38, 526–536. [Google Scholar] [CrossRef]
  30. Lee, M.J.; Wu, Y.; Fried, S.K. Adipose tissue heterogeneity: Implication of depot differences in adipose tissue for obesity complications. Mol. Asp. Med. 2013, 34, 1–11. [Google Scholar] [CrossRef]
  31. Macotela, Y.; Emanuelli, B.; Mori, M.A.; Gesta, S.; Schulz, T.J.; Tseng, Y.H.; Kahn, C.R. Intrinsic differences in adipocyte precursor cells from different white fat depots. Diabetes 2012, 61, 1691–1699. [Google Scholar] [CrossRef] [PubMed]
  32. Morgan-Bathke, M.; Chen, L.; Oberschneider, E.; Harteneck, D.; Jensen, M.D. Sex and depot differences in ex vivo adipose tissue fatty acid storage and glycerol-3-phosphate acyltransferase activity. Am. J. Physiol. Endocrinol. Metab. 2015, 308, E830–E846. [Google Scholar] [CrossRef]
  33. Lee, K.Y.; Luong, Q.; Sharma, R.; Dreyfuss, J.M.; Ussar, S.; Kahn, C.R. Developmental and functional heterogeneity of white adipocytes within a single fat depot. EMBO J. 2019, 38, e99291. [Google Scholar] [CrossRef]
  34. Goossens, G.H.; Jocken, J.W.E.; Blaak, E.E. Sexual dimorphism in cardiometabolic health: The role of adipose tissue, muscle and liver. Nat. Rev. Endocrinol. 2021, 17, 47–66. [Google Scholar] [CrossRef]
  35. Norreen-Thorsen, M.; Struck, E.C.; Oling, S.; Zwahlen, M.; Von Feilitzen, K.; Odeberg, J.; Lindskog, C.; Ponten, F.; Uhlen, M.; Dusart, P.J.; et al. A human adipose tissue cell-type transcriptome atlas. Cell Rep. 2022, 40, 111046. [Google Scholar] [CrossRef]
  36. Bradford, S.T.; Nair, S.S.; Statham, A.L.; van Dijk, S.J.; Peters, T.J.; Anwar, F.; French, H.J.; von Martels, J.Z.H.; Sutcliffe, B.; Maddugoda, M.P.; et al. Methylome and transcriptome maps of human visceral and subcutaneous adipocytes reveal key epigenetic differences at developmental genes. Sci. Rep. 2019, 9, 9511. [Google Scholar] [CrossRef] [PubMed]
  37. Ibrahim, M.M. Subcutaneous and visceral adipose tissue: Structural and functional differences. Obes. Rev. 2010, 11, 11–18. [Google Scholar] [CrossRef] [PubMed]
  38. Sabin, M.A.; Crowne, E.C.; Stewart, C.E.; Hunt, L.P.; Turner, S.J.; Welsh, G.I.; Grohmann, M.J.; Holly, J.M.; Shield, J.P. Depot-specific effects of fatty acids on lipid accumulation in children’s adipocytes. Biochem. Biophys. Res. Commun. 2007, 361, 356–361. [Google Scholar] [CrossRef]
  39. Piers, L.S.; Walker, K.Z.; Stoney, R.M.; Soares, M.J.; O’Dea, K. Substitution of saturated with monounsaturated fat in a 4-week diet affects body weight and composition of overweight and obese men. Br. J. Nutr. 2003, 90, 717–727. [Google Scholar] [CrossRef]
  40. Paniagua, J.A.; Gallego de la Sacristana, A.; Romero, I.; Vidal-Puig, A.; Latre, J.M.; Sanchez, E.; Perez-Martinez, P.; Lopez-Miranda, J.; Perez-Jimenez, F. Monounsaturated fat-rich diet prevents central body fat distribution and decreases postprandial adiponectin expression induced by a carbohydrate-rich diet in insulin-resistant subjects. Diabetes Care 2007, 30, 1717–1723. [Google Scholar] [CrossRef]
  41. Finucane, O.M.; Lyons, C.L.; Murphy, A.M.; Reynolds, C.M.; Klinger, R.; Healy, N.P.; Cooke, A.A.; Coll, R.C.; McAllan, L.; Nilaweera, K.N. Monounsaturated fatty acid–enriched high-fat diets impede adipose NLRP3 inflammasome–mediated IL-1β secretion and insulin resistance despite obesity. Diabetes 2015, 64, 2116–2128. [Google Scholar] [CrossRef] [PubMed]
  42. Gustafson, B.; Hedjazifar, S.; Gogg, S.; Hammarstedt, A.; Smith, U. Insulin resistance and impaired adipogenesis. Trends Endocrinol. Metab. 2015, 26, 193–200. [Google Scholar] [CrossRef]
  43. Hardy, O.T.; Perugini, R.A.; Nicoloro, S.M.; Gallagher-Dorval, K.; Puri, V.; Straubhaar, J.; Czech, M.P. Body mass index-independent inflammation in omental adipose tissue associated with insulin resistance in morbid obesity. Surg. Obes. Relat. Dis. 2011, 7, 60–67. [Google Scholar] [CrossRef] [PubMed]
  44. Kloting, N.; Fasshauer, M.; Dietrich, A.; Kovacs, P.; Schon, M.R.; Kern, M.; Stumvoll, M.; Bluher, M. Insulin-sensitive obesity. Am. J. Physiol. Endocrinol. Metab. 2010, 299, E506–E515. [Google Scholar] [CrossRef]
  45. McLaughlin, T.; Sherman, A.; Tsao, P.; Gonzalez, O.; Yee, G.; Lamendola, C.; Reaven, G.M.; Cushman, S.W. Enhanced proportion of small adipose cells in insulin-resistant vs insulin-sensitive obese individuals implicates impaired adipogenesis. Diabetologia 2007, 50, 1707–1715. [Google Scholar] [CrossRef]
  46. McLaughlin, T.; Lamendola, C.; Coghlan, N.; Liu, T.C.; Lerner, K.; Sherman, A.; Cushman, S.W. Subcutaneous adipose cell size and distribution: Relationship to insulin resistance and body fat. Obesity 2014, 22, 673–680. [Google Scholar] [CrossRef]
  47. Hammarstedt, A.; Gogg, S.; Hedjazifar, S.; Nerstedt, A.; Smith, U. Impaired Adipogenesis and Dysfunctional Adipose Tissue in Human Hypertrophic Obesity. Physiol. Rev. 2018, 98, 1911–1941. [Google Scholar] [CrossRef] [PubMed]
  48. Meln, I.; Wolff, G.; Gajek, T.; Koddebusch, J.; Lerch, S.; Harbrecht, L.; Hong, W.; Bayindir-Buchhalter, I.; Krunic, D.; Augustin, H.G.; et al. Dietary calories and lipids synergistically shape adipose tissue cellularity during postnatal growth. Mol. Metab. 2019, 24, 139–148. [Google Scholar] [CrossRef] [PubMed]
  49. Jo, J.; Gavrilova, O.; Pack, S.; Jou, W.; Mullen, S.; Sumner, A.E.; Cushman, S.W.; Periwal, V. Hypertrophy and/or Hyperplasia: Dynamics of Adipose Tissue Growth. PLoS Comput. Biol. 2009, 5, e1000324. [Google Scholar] [CrossRef]
  50. Weisberg, S.P.; McCann, D.; Desai, M.; Rosenbaum, M.; Leibel, R.L.; Ferrante, A.W., Jr. Obesity is associated with macrophage accumulation in adipose tissue. J. Clin. Investig. 2003, 112, 1796–1808. [Google Scholar] [CrossRef]
  51. Han, C.Y.; Subramanian, S.; Chan, C.K.; Omer, M.; Chiba, T.; Wight, T.N.; Chait, A. Adipocyte-derived serum amyloid A3 and hyaluronan play a role in monocyte recruitment and adhesion. Diabetes 2007, 56, 2260–2273. [Google Scholar] [CrossRef] [PubMed]
  52. Lu, J.; Zhao, J.; Meng, H.; Zhang, X. Adipose Tissue-Resident Immune Cells in Obesity and Type 2 Diabetes. Front. Immunol. 2019, 10, 1173. [Google Scholar] [CrossRef]
  53. Liang, W.; Qi, Y.; Yi, H.; Mao, C.; Meng, Q.; Wang, H.; Zheng, C. The Roles of Adipose Tissue Macrophages in Human Disease. Front. Immunol. 2022, 13, 908749. [Google Scholar] [CrossRef]
  54. O’Rourke, R.W.; White, A.E.; Metcalf, M.D.; Olivas, A.S.; Mitra, P.; Larison, W.G.; Cheang, E.C.; Varlamov, O.; Corless, C.L.; Roberts, C.T., Jr.; et al. Hypoxia-induced inflammatory cytokine secretion in human adipose tissue stromovascular cells. Diabetologia 2011, 54, 1480–1490. [Google Scholar] [CrossRef] [PubMed]
  55. Fujisaka, S.; Usui, I.; Ikutani, M.; Aminuddin, A.; Takikawa, A.; Tsuneyama, K.; Mahmood, A.; Goda, N.; Nagai, Y.; Takatsu, K.; et al. Adipose tissue hypoxia induces inflammatory M1 polarity of macrophages in an HIF-1alpha-dependent and HIF-1alpha-independent manner in obese mice. Diabetologia 2013, 56, 1403–1412. [Google Scholar] [CrossRef] [PubMed]
  56. Xia, W.; Lu, Z.; Chen, W.; Zhou, J.; Zhao, Y. Excess fatty acids induce pancreatic acinar cell pyroptosis through macrophage M1 polarization. BMC Gastroenterol. 2022, 22, 72. [Google Scholar] [CrossRef]
  57. Tiwari, P.; Blank, A.; Cui, C.; Schoenfelt, K.Q.; Zhou, G.; Xu, Y.; Khramtsova, G.; Olopade, F.; Shah, A.M.; Khan, S.A.; et al. Metabolically activated adipose tissue macrophages link obesity to triple-negative breast cancer. J. Exp. Med. 2019, 216, 1345–1358. [Google Scholar] [CrossRef]
  58. Mayi, T.H.; Daoudi, M.; Derudas, B.; Gross, B.; Bories, G.; Wouters, K.; Brozek, J.; Caiazzo, R.; Raverdi, V.; Pigeyre, M.; et al. Human adipose tissue macrophages display activation of cancer-related pathways. J. Biol. Chem. 2012, 287, 21904–21913. [Google Scholar] [CrossRef]
  59. Feuerer, M.; Herrero, L.; Cipolletta, D.; Naaz, A.; Wong, J.; Nayer, A.; Lee, J.; Goldfine, A.B.; Benoist, C.; Shoelson, S.; et al. Lean, but not obese, fat is enriched for a unique population of regulatory T cells that affect metabolic parameters. Nat. Med. 2009, 15, 930–939. [Google Scholar] [CrossRef]
  60. Dyck, L.; Prendeville, H.; Raverdeau, M.; Wilk, M.M.; Loftus, R.M.; Douglas, A.; McCormack, J.; Moran, B.; Wilkinson, M.; Mills, E.L.; et al. Suppressive effects of the obese tumor microenvironment on CD8 T cell infiltration and effector function. J. Exp. Med. 2022, 219, e20210042. [Google Scholar] [CrossRef]
  61. Howie, D.; Ten Bokum, A.; Necula, A.S.; Cobbold, S.P.; Waldmann, H. The Role of Lipid Metabolism in T Lymphocyte Differentiation and Survival. Front. Immunol. 2017, 8, 1949. [Google Scholar] [CrossRef] [PubMed]
  62. Rezalotfi, A.; Ahmadian, E.; Aazami, H.; Solgi, G.; Ebrahimi, M. Gastric Cancer Stem Cells Effect on Th17/Treg Balance; A Bench to Beside Perspective. Front. Oncol. 2019, 9, 226. [Google Scholar] [CrossRef]
  63. Wu, L.; Van Kaer, L. Contribution of lipid-reactive natural killer T cells to obesity-associated inflammation and insulin resistance. Adipocyte 2013, 2, 12–16. [Google Scholar] [CrossRef]
  64. Lynch, L.; O’Shea, D.; Winter, D.C.; Geoghegan, J.; Doherty, D.G.; O’Farrelly, C. Invariant NKT cells and CD1d(+) cells amass in human omentum and are depleted in patients with cancer and obesity. Eur. J. Immunol. 2009, 39, 1893–1901. [Google Scholar] [CrossRef] [PubMed]
  65. Fu, S.; He, K.; Tian, C.; Sun, H.; Zhu, C.; Bai, S.; Liu, J.; Wu, Q.; Xie, D.; Yue, T.; et al. Impaired lipid biosynthesis hinders anti-tumor efficacy of intratumoral iNKT cells. Nat. Commun. 2020, 11, 438. [Google Scholar] [CrossRef] [PubMed]
  66. Duffaut, C.; Galitzky, J.; Lafontan, M.; Bouloumie, A. Unexpected trafficking of immune cells within the adipose tissue during the onset of obesity. Biochem. Biophys. Res. Commun. 2009, 384, 482–485. [Google Scholar] [CrossRef]
  67. Baier, P.K.; Eggstein, S.; Wolff-Vorbeck, G.; Baumgartner, U.; Hopt, U.T. Chemokines in human colorectal carcinoma. Anticancer. Res. 2005, 25, 3581–3584. [Google Scholar] [PubMed]
  68. Waldner, M.J.; Foersch, S.; Neurath, M.F. Interleukin-6—A key regulator of colorectal cancer development. Int. J. Biol. Sci. 2012, 8, 1248–1253. [Google Scholar] [CrossRef]
  69. Winer, D.A.; Winer, S.; Shen, L.; Wadia, P.P.; Yantha, J.; Paltser, G.; Tsui, H.; Wu, P.; Davidson, M.G.; Alonso, M.N.; et al. B cells promote insulin resistance through modulation of T cells and production of pathogenic IgG antibodies. Nat. Med. 2011, 17, 610–617. [Google Scholar] [CrossRef]
  70. DeFuria, J.; Belkina, A.C.; Jagannathan-Bogdan, M.; Snyder-Cappione, J.; Carr, J.D.; Nersesova, Y.R.; Markham, D.; Strissel, K.J.; Watkins, A.A.; Zhu, M.; et al. B cells promote inflammation in obesity and type 2 diabetes through regulation of T-cell function and an inflammatory cytokine profile. Proc. Natl. Acad. Sci. USA 2013, 110, 5133–5138. [Google Scholar] [CrossRef]
  71. Jagannathan, M.; McDonnell, M.; Liang, Y.; Hasturk, H.; Hetzel, J.; Rubin, D.; Kantarci, A.; Van Dyke, T.E.; Ganley-Leal, L.M.; Nikolajczyk, B.S. Toll-like receptors regulate B cell cytokine production in patients with diabetes. Diabetologia 2010, 53, 1461–1471. [Google Scholar] [CrossRef]
  72. Zhang, E.; Ding, C.; Li, S.; Zhou, X.; Aikemu, B.; Fan, X.; Sun, J.; Zheng, M.; Yang, X. Roles and mechanisms of tumour-infiltrating B cells in human cancer: A new force in immunotherapy. Biomark. Res. 2023, 11, 28. [Google Scholar] [CrossRef] [PubMed]
  73. Barr, T.A.; Brown, S.; Mastroeni, P.; Gray, D. TLR and B cell receptor signals to B cells differentially program primary and memory Th1 responses to Salmonella enterica. J. Immunol. 2010, 185, 2783–2789. [Google Scholar] [CrossRef]
  74. Mirlekar, B.; Michaud, D.; Lee, S.J.; Kren, N.P.; Harris, C.; Greene, K.; Goldman, E.C.; Gupta, G.P.; Fields, R.C.; Hawkins, W.G.; et al. B cell-Derived IL35 Drives STAT3-Dependent CD8(+) T-cell Exclusion in Pancreatic Cancer. Cancer Immunol. Res. 2020, 8, 292–308. [Google Scholar] [CrossRef]
  75. Sundara Rajan, S.; Longhi, M.P. Dendritic cells and adipose tissue. Immunology 2016, 149, 353–361. [Google Scholar] [CrossRef] [PubMed]
  76. Herber, D.L.; Cao, W.; Nefedova, Y.; Novitskiy, S.V.; Nagaraj, S.; Tyurin, V.A.; Corzo, A.; Cho, H.I.; Celis, E.; Lennox, B.; et al. Lipid accumulation and dendritic cell dysfunction in cancer. Nat. Med. 2010, 16, 880–886. [Google Scholar] [CrossRef] [PubMed]
  77. James, B.R.; Tomanek-Chalkley, A.; Askeland, E.J.; Kucaba, T.; Griffith, T.S.; Norian, L.A. Diet-induced obesity alters dendritic cell function in the presence and absence of tumor growth. J. Immunol. 2012, 189, 1311–1321. [Google Scholar] [CrossRef]
  78. Xia, S.; Sha, H.; Yang, L.; Ji, Y.; Ostrand-Rosenberg, S.; Qi, L. Gr-1+ CD11b+ myeloid-derived suppressor cells suppress inflammation and promote insulin sensitivity in obesity. J. Biol. Chem. 2011, 286, 23591–23599. [Google Scholar] [CrossRef]
  79. Bao, Y.; Mo, J.; Ruan, L.; Li, G. Increased monocytic CD14(+)HLADRlow/-myeloid-derived suppressor cells in obesity. Mol. Med. Rep. 2015, 11, 2322–2328. [Google Scholar] [CrossRef]
  80. Yan, D.; Adeshakin, A.O.; Xu, M.; Afolabi, L.O.; Zhang, G.; Chen, Y.H.; Wan, X. Lipid Metabolic Pathways Confer the Immunosuppressive Function of Myeloid-Derived Suppressor Cells in Tumor. Front. Immunol. 2019, 10, 1399. [Google Scholar] [CrossRef]
  81. Clements, V.K.; Long, T.; Long, R.; Figley, C.; Smith, D.M.C.; Ostrand-Rosenberg, S. Frontline Science: High fat diet and leptin promote tumor progression by inducing myeloid-derived suppressor cells. J. Leukoc. Biol. 2018, 103, 395–407. [Google Scholar] [CrossRef] [PubMed]
  82. Sanchez-Pino, M.D.; Richardson, W.S.; Zabaleta, J.; Puttalingaiah, R.T.; Chapple, A.G.; Liu, J.; Kim, Y.; Ponder, M.; DeArmitt, R.; Baiamonte, L.B.; et al. Increased inflammatory low-density neutrophils in severe obesity and effect of bariatric surgery: Results from case-control and prospective cohort studies. eBioMedicine 2022, 77, 103910. [Google Scholar] [CrossRef]
  83. Rice, C.M.; Davies, L.C.; Subleski, J.J.; Maio, N.; Gonzalez-Cotto, M.; Andrews, C.; Patel, N.L.; Palmieri, E.M.; Weiss, J.M.; Lee, J.M.; et al. Tumour-elicited neutrophils engage mitochondrial metabolism to circumvent nutrient limitations and maintain immune suppression. Nat. Commun. 2018, 9, 5099. [Google Scholar] [CrossRef]
  84. Xiong, S.; Dong, L.; Cheng, L. Neutrophils in cancer carcinogenesis and metastasis. J. Hematol. Oncol. 2021, 14, 173. [Google Scholar] [CrossRef] [PubMed]
  85. Quail, D.F.; Olson, O.C.; Bhardwaj, P.; Walsh, L.A.; Akkari, L.; Quick, M.L.; Chen, I.C.; Wendel, N.; Ben-Chetrit, N.; Walker, J.; et al. Obesity alters the lung myeloid cell landscape to enhance breast cancer metastasis through IL5 and GM-CSF. Nat. Cell Biol. 2017, 19, 974–987. [Google Scholar] [CrossRef]
  86. Lynch, L.A.; O’Connell, J.M.; Kwasnik, A.K.; Cawood, T.J.; O’Farrelly, C.; O’Shea, D.B. Are natural killer cells protecting the metabolically healthy obese patient? Obesity 2009, 17, 601–605. [Google Scholar] [CrossRef] [PubMed]
  87. Kobayashi, T.; Lam, P.Y.; Jiang, H.; Bednarska, K.; Gloury, R.; Murigneux, V.; Tay, J.; Jacquelot, N.; Li, R.; Tuong, Z.K.; et al. Increased lipid metabolism impairs NK cell function and mediates adaptation to the lymphoma environment. Blood 2020, 136, 3004–3017. [Google Scholar] [CrossRef]
  88. Conroy, M.J.; Fitzgerald, V.; Doyle, S.L.; Channon, S.; Useckaite, Z.; Gilmartin, N.; O’Farrelly, C.; Ravi, N.; Reynolds, J.V.; Lysaght, J. The microenvironment of visceral adipose tissue and liver alter natural killer cell viability and function. J. Leukoc. Biol. 2016, 100, 1435–1442. [Google Scholar] [CrossRef]
  89. Mylod, E.; O’Connell, F.; Donlon, N.E.; Butler, C.; Reynolds, J.V.; Lysaght, J.; Conroy, M.J. The Omentum in Obesity-Associated Cancer: A Hindrance to Effective Natural Killer Cell Migration towards Tumour Which Can Be Overcome by CX3CR1 Antagonism. Cancers 2021, 14, 64. [Google Scholar] [CrossRef]
  90. Ferrante, A.W., Jr. The immune cells in adipose tissue. Diabetes Obes. Metab. 2013, 15 (Suppl. S3), 34–38. [Google Scholar] [CrossRef]
  91. McGillicuddy, F.C.; Harford, K.A.; Reynolds, C.M.; Oliver, E.; Claessens, M.; Mills, K.H.; Roche, H.M. Lack of interleukin-1 receptor I (IL-1RI) protects mice from high-fat diet-induced adipose tissue inflammation coincident with improved glucose homeostasis. Diabetes 2011, 60, 1688–1698. [Google Scholar] [CrossRef] [PubMed]
  92. Reynolds, C.M.; McGillicuddy, F.C.; Harford, K.A.; Finucane, O.M.; Mills, K.H.; Roche, H.M. Dietary saturated fatty acids prime the NLRP3 inflammasome via TLR4 in dendritic cells-implications for diet-induced insulin resistance. Mol. Nutr. Food Res. 2012, 56, 1212–1222. [Google Scholar] [CrossRef] [PubMed]
  93. Wen, H.; Gris, D.; Lei, Y.; Jha, S.; Zhang, L.; Huang, M.T.; Brickey, W.J.; Ting, J.P. Fatty acid-induced NLRP3-ASC inflammasome activation interferes with insulin signaling. Nat. Immunol. 2011, 12, 408–415. [Google Scholar] [CrossRef] [PubMed]
  94. Youm, Y.H.; Adijiang, A.; Vandanmagsar, B.; Burk, D.; Ravussin, A.; Dixit, V.D. Elimination of the NLRP3-ASC inflammasome protects against chronic obesity-induced pancreatic damage. Endocrinology 2011, 152, 4039–4045. [Google Scholar] [CrossRef] [PubMed]
  95. Bradley, R.L.; Fisher, F.F.; Maratos-Flier, E. Dietary fatty acids differentially regulate production of TNF-alpha and IL-10 by murine 3T3-L1 adipocytes. Obesity 2008, 16, 938–944. [Google Scholar] [CrossRef]
  96. Chan, K.L.; Pillon, N.J.; Sivaloganathan, D.M.; Costford, S.R.; Liu, Z.; Theret, M.; Chazaud, B.; Klip, A. Palmitoleate Reverses High Fat-induced Proinflammatory Macrophage Polarization via AMP-activated Protein Kinase (AMPK). J. Biol. Chem. 2015, 290, 16979–16988. [Google Scholar] [CrossRef]
  97. Zhou, B.R.; Zhang, J.A.; Zhang, Q.; Permatasari, F.; Xu, Y.; Wu, D.; Yin, Z.Q.; Luo, D. Palmitic acid induces production of proinflammatory cytokines interleukin-6, interleukin-1beta, and tumor necrosis factor-alpha via a NF-kappaB-dependent mechanism in HaCaT keratinocytes. Mediat. Inflamm. 2013, 2013, 530429. [Google Scholar] [CrossRef]
  98. Korbecki, J.; Bajdak-Rusinek, K. The effect of palmitic acid on inflammatory response in macrophages: An overview of molecular mechanisms. Inflamm. Res. 2019, 68, 915–932. [Google Scholar] [CrossRef]
  99. De Lima, T.M.; de Sa Lima, L.; Scavone, C.; Curi, R. Fatty acid control of nitric oxide production by macrophages. FEBS Lett. 2006, 580, 3287–3295. [Google Scholar] [CrossRef]
  100. Lancaster, G.I.; Langley, K.G.; Berglund, N.A.; Kammoun, H.L.; Reibe, S.; Estevez, E.; Weir, J.; Mellett, N.A.; Pernes, G.; Conway, J.R.W.; et al. Evidence that TLR4 Is Not a Receptor for Saturated Fatty Acids but Mediates Lipid-Induced Inflammation by Reprogramming Macrophage Metabolism. Cell Metab. 2018, 27, 1096–1110.e1095. [Google Scholar] [CrossRef]
  101. Harvey, K.A.; Walker, C.L.; Xu, Z.; Whitley, P.; Pavlina, T.M.; Hise, M.; Zaloga, G.P.; Siddiqui, R.A. Oleic acid inhibits stearic acid-induced inhibition of cell growth and pro-inflammatory responses in human aortic endothelial cells. J. Lipid Res. 2010, 51, 3470–3480. [Google Scholar] [CrossRef]
  102. Camell, C.; Smith, C.W. Dietary oleic acid increases m2 macrophages in the mesenteric adipose tissue. PLoS ONE 2013, 8, e75147. [Google Scholar] [CrossRef] [PubMed]
  103. Stentz, F.B.; Kitabchi, A.E. Palmitic acid-induced activation of human T-lymphocytes and aortic endothelial cells with production of insulin receptors, reactive oxygen species, cytokines, and lipid peroxidation. Biochem. Biophys. Res. Commun. 2006, 346, 721–726. [Google Scholar] [CrossRef] [PubMed]
  104. Takahashi, H.K.; Cambiaghi, T.D.; Luchessi, A.D.; Hirabara, S.M.; Vinolo, M.A.; Newsholme, P.; Curi, R. Activation of survival and apoptotic signaling pathways in lymphocytes exposed to palmitic acid. J. Cell. Physiol. 2012, 227, 339–350. [Google Scholar] [CrossRef]
  105. Verlengia, R.; Gorjao, R.; Kanunfre, C.C.; Bordin, S.; de Lima, T.M.; Curi, R. Effect of arachidonic acid on proliferation, cytokines production and pleiotropic genes expression in Jurkat cells--a comparison with oleic acid. Life Sci. 2003, 73, 2939–2951. [Google Scholar] [CrossRef]
  106. Frasca, D.; Romero, M.; Garcia, D.; Diaz, A.; Blomberg, B.B. Obesity Accelerates Age-Associated Defects in Human B Cells Through a Metabolic Reprogramming Induced by the Fatty Acid Palmitate. Front. Aging 2021, 2, 828697. [Google Scholar] [CrossRef]
  107. Zhou, X.; Zhu, X.; Li, C.; Li, Y.; Ye, Z.; Shapiro, V.S.; Copland, J.A., 3rd; Hitosugi, T.; Bernlohr, D.A.; Sun, J.; et al. Stearoyl-CoA Desaturase-Mediated Monounsaturated Fatty Acid Availability Supports Humoral Immunity. Cell Rep. 2021, 34, 108601. [Google Scholar] [CrossRef] [PubMed]
  108. Nicholas, D.A.; Zhang, K.; Hung, C.; Glasgow, S.; Aruni, A.W.; Unternaehrer, J.; Payne, K.J.; Langridge, W.H.R.; De Leon, M. Palmitic acid is a toll-like receptor 4 ligand that induces human dendritic cell secretion of IL-1beta. PLoS ONE 2017, 12, e0176793. [Google Scholar] [CrossRef]
  109. Wanten, G.J.; Janssen, F.P.; Naber, A.H. Saturated triglycerides and fatty acids activate neutrophils depending on carbon chain-length. Eur. J. Clin. Investig. 2002, 32, 285–289. [Google Scholar] [CrossRef]
  110. Tam, T.H.; Chan, K.L.; Boroumand, P.; Liu, Z.; Brozinick, J.T.; Bui, H.H.; Roth, K.; Wakefield, C.B.; Penuela, S.; Bilan, P.J.; et al. Nucleotides released from palmitate-activated murine macrophages attract neutrophils. J. Biol. Chem. 2020, 295, 4902–4911. [Google Scholar] [CrossRef]
  111. Hatanaka, E.; Levada-Pires, A.C.; Pithon-Curi, T.C.; Curi, R. Systematic study on ROS production induced by oleic, linoleic, and gamma-linolenic acids in human and rat neutrophils. Free Radic. Biol. Med. 2006, 41, 1124–1132. [Google Scholar] [CrossRef]
  112. Medeiros-de-Moraes, I.M.; Goncalves-de-Albuquerque, C.F.; Kurz, A.R.M.; Oliveira, F.M.J.; de Abreu, V.H.P.; Torres, R.C.; Carvalho, V.F.; Estato, V.; Bozza, P.T.; Sperandio, M.; et al. Omega-9 Oleic Acid, the Main Compound of Olive Oil, Mitigates Inflammation during Experimental Sepsis. Oxid. Med. Cell. Longev. 2018, 2018, 6053492. [Google Scholar] [CrossRef] [PubMed]
  113. Hidalgo, M.A.; Nahuelpan, C.; Manosalva, C.; Jara, E.; Carretta, M.D.; Conejeros, I.; Loaiza, A.; Chihuailaf, R.; Burgos, R.A. Oleic acid induces intracellular calcium mobilization, MAPK phosphorylation, superoxide production and granule release in bovine neutrophils. Biochem. Biophys. Res. Commun. 2011, 409, 280–286. [Google Scholar] [CrossRef] [PubMed]
  114. Pereira, L.M.; Hatanaka, E.; Martins, E.F.; Oliveira, F.; Liberti, E.A.; Farsky, S.H.; Curi, R.; Pithon-Curi, T.C. Effect of oleic and linoleic acids on the inflammatory phase of wound healing in rats. Cell Biochem. Funct. 2008, 26, 197–204. [Google Scholar] [CrossRef] [PubMed]
  115. Alarcon, P.; Manosalva, C.; Quiroga, J.; Belmar, I.; Alvarez, K.; Diaz, G.; Taubert, A.; Hermosilla, C.; Carretta, M.D.; Burgos, R.A.; et al. Oleic and Linoleic Acids Induce the Release of Neutrophil Extracellular Traps via Pannexin 1-Dependent ATP Release and P2X1 Receptor Activation. Front. Vet. Sci. 2020, 7, 260. [Google Scholar] [CrossRef]
  116. DeBerardinis, R.J.; Thompson, C.B. Cellular metabolism and disease: What do metabolic outliers teach us? Cell 2012, 148, 1132–1144. [Google Scholar] [CrossRef]
  117. Hsu, C.C.; Tseng, L.M.; Lee, H.C. Role of mitochondrial dysfunction in cancer progression. Exp. Biol. Med. 2016, 241, 1281–1295. [Google Scholar] [CrossRef]
  118. Desbats, M.A.; Giacomini, I.; Prayer-Galetti, T.; Montopoli, M. Metabolic Plasticity in Chemotherapy Resistance. Front. Oncol. 2020, 10, 281. [Google Scholar] [CrossRef]
  119. Shimura, T.; Noma, N.; Sano, Y.; Ochiai, Y.; Oikawa, T.; Fukumoto, M.; Kunugita, N. AKT-mediated enhanced aerobic glycolysis causes acquired radioresistance by human tumor cells. Radiother. Oncol. 2014, 112, 302–307. [Google Scholar] [CrossRef]
  120. Germain, N.; Dhayer, M.; Boileau, M.; Fovez, Q.; Kluza, J.; Marchetti, P. Lipid Metabolism and Resistance to Anticancer Treatment. Biology 2020, 9, 474. [Google Scholar] [CrossRef]
  121. Butler, L.M.; Perone, Y.; Dehairs, J.; Lupien, L.E.; de Laat, V.; Talebi, A.; Loda, M.; Kinlaw, W.B.; Swinnen, J.V. Lipids and cancer: Emerging roles in pathogenesis, diagnosis and therapeutic intervention. Adv. Drug Deliv. Rev. 2020, 159, 245–293. [Google Scholar] [CrossRef]
  122. She, K.; Fang, S.; Du, W.; Fan, X.; He, J.; Pan, H.; Huang, L.; He, P.; Huang, J. SCD1 is required for EGFR-targeting cancer therapy of lung cancer via re-activation of EGFR/PI3K/AKT signals. Cancer Cell Int. 2019, 19, 103. [Google Scholar] [CrossRef]
  123. Aloia, A.; Mullhaupt, D.; Chabbert, C.D.; Eberhart, T.; Fluckiger-Mangual, S.; Vukolic, A.; Eichhoff, O.; Irmisch, A.; Alexander, L.T.; Scibona, E.; et al. A Fatty Acid Oxidation-dependent Metabolic Shift Regulates the Adaptation of BRAF-mutated Melanoma to MAPK Inhibitors. Clin. Cancer Res. 2019, 25, 6852–6867. [Google Scholar] [CrossRef]
  124. Luo, X.; Cheng, C.; Tan, Z.; Li, N.; Tang, M.; Yang, L.; Cao, Y. Emerging roles of lipid metabolism in cancer metastasis. Mol. Cancer 2017, 16, 76. [Google Scholar] [CrossRef]
  125. Batista-Gonzalez, A.; Vidal, R.; Criollo, A.; Carreno, L.J. New Insights on the Role of Lipid Metabolism in the Metabolic Reprogramming of Macrophages. Front. Immunol. 2019, 10, 2993. [Google Scholar] [CrossRef]
  126. O’Sullivan, D.; van der Windt, G.J.; Huang, S.C.; Curtis, J.D.; Chang, C.H.; Buck, M.D.; Qiu, J.; Smith, A.M.; Lam, W.Y.; DiPlato, L.M.; et al. Memory CD8+ T cells use cell-intrinsic lipolysis to support the metabolic programming necessary for development. Immunity 2014, 41, 75–88. [Google Scholar] [CrossRef]
  127. Kempkes, R.W.M.; Joosten, I.; Koenen, H.; He, X. Metabolic Pathways Involved in Regulatory T Cell Functionality. Front. Immunol. 2019, 10, 2839. [Google Scholar] [CrossRef] [PubMed]
  128. Koundouros, N.; Poulogiannis, G. Reprogramming of fatty acid metabolism in cancer. Br. J. Cancer 2020, 122, 4–22. [Google Scholar] [CrossRef]
  129. Mayas, M.D.; Ortega, F.J.; Macias-Gonzalez, M.; Bernal, R.; Gomez-Huelgas, R.; Fernandez-Real, J.M.; Tinahones, F.J. Inverse relation between FASN expression in human adipose tissue and the insulin resistance level. Nutr. Metab. 2010, 7, 3. [Google Scholar] [CrossRef] [PubMed]
  130. Mobbs, C.V.; Makimura, H. Block the FAS, lose the fat. Nat. Med. 2002, 8, 335–336. [Google Scholar] [CrossRef] [PubMed]
  131. Khasawneh, J.; Schulz, M.D.; Walch, A.; Rozman, J.; Hrabe de Angelis, M.; Klingenspor, M.; Buck, A.; Schwaiger, M.; Saur, D.; Schmid, R.M.; et al. Inflammation and mitochondrial fatty acid beta-oxidation link obesity to early tumor promotion. Proc. Natl. Acad. Sci. USA 2009, 106, 3354–3359. [Google Scholar] [CrossRef]
  132. Menendez, J.A.; Lupu, R. Fatty acid synthase regulates estrogen receptor-alpha signaling in breast cancer cells. Oncogenesis 2017, 6, e299. [Google Scholar] [CrossRef] [PubMed]
  133. Carracedo, A.; Cantley, L.C.; Pandolfi, P.P. Cancer metabolism: Fatty acid oxidation in the limelight. Nat. Rev. Cancer 2013, 13, 227–232. [Google Scholar] [CrossRef] [PubMed]
  134. Bhatt, A.N.; Chauhan, A.; Khanna, S.; Rai, Y.; Singh, S.; Soni, R.; Kalra, N.; Dwarakanath, B.S. Transient elevation of glycolysis confers radio-resistance by facilitating DNA repair in cells. BMC Cancer 2015, 15, 335. [Google Scholar] [CrossRef]
  135. Li, M.; Xian, H.C.; Tang, Y.J.; Liang, X.H.; Tang, Y.L. Fatty acid oxidation: Driver of lymph node metastasis. Cancer Cell Int. 2021, 21, 339. [Google Scholar] [CrossRef]
  136. Hu, Y.; Xu, W.; Zeng, H.; He, Z.; Lu, X.; Zuo, D.; Qin, G.; Chen, W. OXPHOS-dependent metabolic reprogramming prompts metastatic potential of breast cancer cells under osteogenic differentiation. Br. J. Cancer 2020, 123, 1644–1655. [Google Scholar] [CrossRef]
  137. Parlani, M.; Jorgez, C.; Friedl, P. Plasticity of cancer invasion and energy metabolism. Trends Cell Biol. 2023, 33, 388–402. [Google Scholar] [CrossRef] [PubMed]
  138. Massague, J.; Obenauf, A.C. Metastatic colonization by circulating tumour cells. Nature 2016, 529, 298–306. [Google Scholar] [CrossRef]
  139. Gerber, S.A.; Rybalko, V.Y.; Bigelow, C.E.; Lugade, A.A.; Foster, T.H.; Frelinger, J.G.; Lord, E.M. Preferential attachment of peritoneal tumor metastases to omental immune aggregates and possible role of a unique vascular microenvironment in metastatic survival and growth. Am. J. Pathol. 2006, 169, 1739–1752. [Google Scholar] [CrossRef]
  140. Herold, J.; Kalucka, J. Angiogenesis in Adipose Tissue: The Interplay Between Adipose and Endothelial Cells. Front. Physiol. 2020, 11, 624903. [Google Scholar] [CrossRef]
  141. O’Connell, F.; O’Sullivan, J. Help or hindrance: The obesity paradox in cancer treatment response. Cancer Lett. 2021, 522, 269–280. [Google Scholar] [CrossRef] [PubMed]
  142. Deng, T.; Lyon, C.J.; Bergin, S.; Caligiuri, M.A.; Hsueh, W.A. Obesity, Inflammation, and Cancer. Annu. Rev. Pathol. 2016, 11, 421–449. [Google Scholar] [CrossRef] [PubMed]
  143. Donohoe, C.L.; Lysaght, J.; O’Sullivan, J.; Reynolds, J.V. Emerging Concepts Linking Obesity with the Hallmarks of Cancer. Trends Endocrinol. Metab. 2017, 28, 46–62. [Google Scholar] [CrossRef] [PubMed]
  144. Medzhitov, R. Origin and physiological roles of inflammation. Nature 2008, 454, 428–435. [Google Scholar] [CrossRef]
  145. Coussens, L.M.; Werb, Z. Inflammation and cancer. Nature 2002, 420, 860–867. [Google Scholar] [CrossRef]
  146. Elinav, E.; Nowarski, R.; Thaiss, C.A.; Hu, B.; Jin, C.; Flavell, R.A. Inflammation-induced cancer: Crosstalk between tumours, immune cells and microorganisms. Nat. Rev. Cancer 2013, 13, 759–771. [Google Scholar] [CrossRef]
  147. Ribeiro, R.J.; Monteiro, C.P.; Cunha, V.F.; Azevedo, A.S.; Oliveira, M.J.; Monteiro, R.; Fraga, A.M.; Principe, P.; Lobato, C.; Lobo, F.; et al. Tumor cell-educated periprostatic adipose tissue acquires an aggressive cancer-promoting secretory profile. Cell Physiol. Biochem. 2012, 29, 233–240. [Google Scholar] [CrossRef]
  148. Colotta, F.; Allavena, P.; Sica, A.; Garlanda, C.; Mantovani, A. Cancer-related inflammation, the seventh hallmark of cancer: Links to genetic instability. Carcinogenesis 2009, 30, 1073–1081. [Google Scholar] [CrossRef]
  149. Liu, R.; Nikolajczyk, B.S. Tissue Immune Cells Fuel Obesity-Associated Inflammation in Adipose Tissue and Beyond. Front. Immunol. 2019, 10, 1587. [Google Scholar] [CrossRef]
  150. Brown, L.M.; Swanson, C.A.; Gridley, G.; Swanson, G.M.; Schoenberg, J.B.; Greenberg, R.S.; Silverman, D.T.; Pottern, L.M.; Hayes, R.B.; Schwartz, A.G.; et al. Adenocarcinoma of the esophagus: Role of obesity and diet. J. Natl. Cancer Inst. 1995, 87, 104–109. [Google Scholar] [CrossRef]
  151. Vaughan, T.L.; Davis, S.; Kristal, A.; Thomas, D.B. Obesity, alcohol, and tobacco as risk factors for cancers of the esophagus and gastric cardia: Adenocarcinoma versus squamous cell carcinoma. Cancer Epidemiol. Biomark. Prev. 1995, 4, 85–92. [Google Scholar]
  152. Ryan, A.M.; Rowley, S.P.; Fitzgerald, A.P.; Ravi, N.; Reynolds, J.V. Adenocarcinoma of the oesophagus and gastric cardia: Male preponderance in association with obesity. Eur. J. Cancer 2006, 42, 1151–1158. [Google Scholar] [CrossRef] [PubMed]
  153. Renehan, A.G.; Roberts, D.L.; Dive, C. Obesity and cancer: Pathophysiological and biological mechanisms. Arch. Physiol. Biochem. 2008, 114, 71–83. [Google Scholar] [CrossRef] [PubMed]
  154. Thrift, A.P.; Shaheen, N.J.; Gammon, M.D.; Bernstein, L.; Reid, B.J.; Onstad, L.; Risch, H.A.; Liu, G.; Bird, N.C.; Wu, A.H.; et al. Obesity and risk of esophageal adenocarcinoma and Barrett’s esophagus: A Mendelian randomization study. J. Natl. Cancer Inst. 2014, 106, dju252. [Google Scholar] [CrossRef] [PubMed]
  155. Conroy, M.J.; Maher, S.G.; Melo, A.M.; Doyle, S.L.; Foley, E.; Reynolds, J.V.; Long, A.; Lysaght, J. Identifying a Novel Role for Fractalkine (CX3CL1) in Memory CD8(+) T Cell Accumulation in the Omentum of Obesity-Associated Cancer Patients. Front. Immunol. 2018, 9, 1867. [Google Scholar] [CrossRef] [PubMed]
  156. Melo, A.M.; Mylod, E.; Fitzgerald, V.; Donlon, N.E.; Murphy, D.M.; Foley, E.K.; Bhardwaj, A.; Reynolds, J.V.; Doherty, D.G.; Lysaght, J.; et al. Tissue distribution of gammadelta T cell subsets in oesophageal adenocarcinoma. Clin. Immunol. 2021, 229, 108797. [Google Scholar] [CrossRef]
  157. Davern, M.; Bracken-Clarke, D.; Donlon, N.E.; Sheppard, A.D.; Connell, F.O.; Heeran, A.B.; Majcher, K.; Conroy, M.J.; Mylod, E.; Butler, C.; et al. Visceral adipose tissue secretome from early and late-stage oesophageal cancer patients differentially affects effector and regulatory T cells. J. Cancer Res. Clin. Oncol. 2023, 149, 6583–6599. [Google Scholar] [CrossRef]
  158. Donlon, N.E.; Sheppard, A.; Davern, M.; O’Connell, F.; Phelan, J.J.; Power, R.; Nugent, T.; Dinneen, K.; Aird, J.; Greene, J.; et al. Linking Circulating Serum Proteins with Clinical Outcomes in Esophageal Adenocarcinoma-An Emerging Role for Chemokines. Cancers 2020, 12, 3356. [Google Scholar] [CrossRef]
  159. Heeran, A.B.; McCready, J.; Dunne, M.R.; Donlon, N.E.; Nugent, T.S.; Bhardwaj, A.; Mitchelson, K.A.J.; Buckley, A.M.; Ravi, N.; Roche, H.M.; et al. Opposing Immune-Metabolic Signature in Visceral Versus Subcutaneous Adipose Tissue in Patients with Adenocarcinoma of the Oesophagus and the Oesophagogastric Junction. Metabolites 2021, 11, 768. [Google Scholar] [CrossRef]
  160. O’Connell, F.; Mylod, E.; Donlon, N.E.; Heeran, A.B.; Butler, C.; Bhardwaj, A.; Ramjit, S.; Durand, M.; Lambe, G.; Tansey, P.; et al. Energy Metabolism, Metabolite, and Inflammatory Profiles in Human Ex Vivo Adipose Tissue Are Influenced by Obesity Status, Metabolic Dysfunction, and Treatment Regimes in Patients with Oesophageal Adenocarcinoma. Cancers 2023, 15, 1681. [Google Scholar] [CrossRef]
  161. Hashimoto, I.; Wada, J.; Hida, A.; Baba, M.; Miyatake, N.; Eguchi, J.; Shikata, K.; Makino, H. Elevated serum monocyte chemoattractant protein-4 and chronic inflammation in overweight subjects. Obesity 2006, 14, 799–811. [Google Scholar] [CrossRef] [PubMed]
  162. Brown, K.F.; Rumgay, H.; Dunlop, C.; Ryan, M.; Quartly, F.; Cox, A.; Deas, A.; Elliss-Brookes, L.; Gavin, A.; Hounsome, L.; et al. The fraction of cancer attributable to modifiable risk factors in England, Wales, Scotland, Northern Ireland, and the United Kingdom in 2015. Br. J. Cancer 2018, 118, 1130–1141. [Google Scholar] [CrossRef]
  163. Kersy, O.; Loewenstein, S.; Lubezky, N.; Sher, O.; Simon, N.B.; Klausner, J.M.; Lahat, G. Omental Tissue-Mediated Tumorigenesis of Gastric Cancer Peritoneal Metastases. Front. Oncol. 2019, 9, 1267. [Google Scholar] [CrossRef]
  164. Hamabe-Horiike, T.; Harada, S.I.; Yoshida, K.; Kinoshita, J.; Yamaguchi, T.; Fushida, S. Adipocytes contribute to tumor progression and invasion of peritoneal metastasis by interacting with gastric cancer cells as cancer associated fibroblasts. Cancer Rep. 2023, 6, e1647. [Google Scholar] [CrossRef] [PubMed]
  165. Natsume, M.; Shimura, T.; Iwasaki, H.; Okuda, Y.; Hayashi, K.; Takahashi, S.; Kataoka, H. Omental adipocytes promote peritoneal metastasis of gastric cancer through the CXCL2-VEGFA axis. Br. J. Cancer 2020, 123, 459–470. [Google Scholar] [CrossRef] [PubMed]
  166. Alkhatatbeh, M.J.; Enjeti, A.K.; Acharya, S.; Thorne, R.F.; Lincz, L.F. The origin of circulating CD36 in type 2 diabetes. Nutr. Diabetes 2013, 3, e59. [Google Scholar] [CrossRef]
  167. Majchrzak, K.; Piotrowska, M.; Krajewska, J.; Fichna, J. Adipocyte Fatty Acid Binding Protein (A-FABP) as a Potential New Therapeutic Target for the Treatment of Obesity—Associated Cancers. Curr. Drug Targets 2022, 23, 597–605. [Google Scholar] [CrossRef]
  168. Singal, A.G.; El-Serag, H.B. Hepatocellular Carcinoma from Epidemiology to Prevention: Translating Knowledge into Practice. Clin. Gastroenterol. Hepatol. 2015, 13, 2140–2151. [Google Scholar] [CrossRef]
  169. Allen, A.M.; Hicks, S.B.; Mara, K.C.; Larson, J.J.; Therneau, T.M. The risk of incident extrahepatic cancers is higher in non-alcoholic fatty liver disease than obesity—A longitudinal cohort study. J. Hepatol. 2019, 71, 1229–1236. [Google Scholar] [CrossRef]
  170. Montano-Loza, A.J.; Mazurak, V.C.; Ebadi, M.; Meza-Junco, J.; Sawyer, M.B.; Baracos, V.E.; Kneteman, N. Visceral adiposity increases risk for hepatocellular carcinoma in male patients with cirrhosis and recurrence after liver transplant. Hepatology 2018, 67, 914–923. [Google Scholar] [CrossRef]
  171. Khoramipour, K.; Chamari, K.; Hekmatikar, A.A.; Ziyaiyan, A.; Taherkhani, S.; Elguindy, N.M.; Bragazzi, N.L. Adiponectin: Structure, Physiological Functions, Role in Diseases, and Effects of Nutrition. Nutrients 2021, 13, 1180. [Google Scholar] [CrossRef] [PubMed]
  172. Zhang, L.; Yuan, Q.; Li, M.; Chai, D.; Deng, W.; Wang, W. The association of leptin and adiponectin with hepatocellular carcinoma risk and prognosis: A combination of traditional, survival, and dose-response meta-analysis. BMC Cancer 2020, 20, 1167. [Google Scholar] [CrossRef] [PubMed]
  173. Park, E.J.; Lee, J.H.; Yu, G.Y.; He, G.; Ali, S.R.; Holzer, R.G.; Osterreicher, C.H.; Takahashi, H.; Karin, M. Dietary and genetic obesity promote liver inflammation and tumorigenesis by enhancing IL-6 and TNF expression. Cell 2010, 140, 197–208. [Google Scholar] [CrossRef] [PubMed]
  174. Larsson, S.C.; Wolk, A. Overweight, obesity and risk of liver cancer: A meta-analysis of cohort studies. Br. J. Cancer 2007, 97, 1005–1008. [Google Scholar] [CrossRef]
  175. Banales, J.M.; Cardinale, V.; Carpino, G.; Marzioni, M.; Andersen, J.B.; Invernizzi, P.; Lind, G.E.; Folseraas, T.; Forbes, S.J.; Fouassier, L.; et al. Expert consensus document: Cholangiocarcinoma: Current knowledge and future perspectives consensus statement from the European Network for the Study of Cholangiocarcinoma (ENS-CCA). Nat. Rev. Gastroenterol. Hepatol. 2016, 13, 261–280. [Google Scholar] [CrossRef]
  176. Yugawa, K.; Itoh, S.; Iseda, N.; Kurihara, T.; Kitamura, Y.; Toshima, T.; Harada, N.; Kohashi, K.; Baba, S.; Ishigami, K.; et al. Obesity is a risk factor for intrahepatic cholangiocarcinoma progression associated with alterations of metabolic activity and immune status. Sci. Rep. 2021, 11, 5845. [Google Scholar] [CrossRef]
  177. Bardou, M.; Rouland, A.; Martel, M.; Loffroy, R.; Barkun, A.N.; Chapelle, N. Review article: Obesity and colorectal cancer. Aliment. Pharmacol. Ther. 2022, 56, 407–418. [Google Scholar] [CrossRef]
  178. Edwards, R.A.; Witherspoon, M.; Wang, K.; Afrasiabi, K.; Pham, T.; Birnbaumer, L.; Lipkin, S.M. Epigenetic repression of DNA mismatch repair by inflammation and hypoxia in inflammatory bowel disease-associated colorectal cancer. Cancer Res. 2009, 69, 6423–6429. [Google Scholar] [CrossRef]
  179. Wunderlich, C.M.; Ackermann, P.J.; Ostermann, A.L.; Adams-Quack, P.; Vogt, M.C.; Tran, M.L.; Nikolajev, A.; Waisman, A.; Garbers, C.; Theurich, S.; et al. Obesity exacerbates colitis-associated cancer via IL-6-regulated macrophage polarisation and CCL-20/CCR-6-mediated lymphocyte recruitment. Nat. Commun. 2018, 9, 1646. [Google Scholar] [CrossRef]
  180. Nanno, M.; Kanari, Y.; Naito, T.; Inoue, N.; Hisamatsu, T.; Chinen, H.; Sugimoto, K.; Shimomura, Y.; Yamagishi, H.; Shiohara, T.; et al. Exacerbating role of gammadelta T cells in chronic colitis of T-cell receptor alpha mutant mice. Gastroenterology 2008, 134, 481–490. [Google Scholar] [CrossRef]
  181. Tie, G.; Yan, J.; Khair, L.; Messina, J.A.; Deng, A.; Kang, J.; Fazzio, T.; Messina, L.M. Hypercholesterolemia Increases Colorectal Cancer Incidence by Reducing Production of NKT and gammadelta T Cells from Hematopoietic Stem Cells. Cancer Res. 2017, 77, 2351–2362. [Google Scholar] [CrossRef] [PubMed]
  182. Del Corno, M.; D’Archivio, M.; Conti, L.; Scazzocchio, B.; Vari, R.; Donninelli, G.; Varano, B.; Giammarioli, S.; De Meo, S.; Silecchia, G.; et al. Visceral fat adipocytes from obese and colorectal cancer subjects exhibit distinct secretory and omega6 polyunsaturated fatty acid profiles and deliver immunosuppressive signals to innate immunity cells. Oncotarget 2016, 7, 63093–63105. [Google Scholar] [CrossRef] [PubMed]
  183. Basile, D.; Rosati, G.; Bergamo, F.; Garattini, S.K.; Banzi, M.; Zampino, M.; Bozzarelli, S.; Marchetti, P.; Galli, F.; Galli, F.; et al. Prognostic Value of Body Mass Index in Stage II/III Colon Cancer: Posthoc Analysis From the TOSCA Trial. Clin. Color. Cancer 2023, 22, 190–198. [Google Scholar] [CrossRef] [PubMed]
  184. Assumpcao, J.A.F.; Pasquarelli-do-Nascimento, G.; Duarte, M.S.V.; Bonamino, M.H.; Magalhaes, K.G. The ambiguous role of obesity in oncology by promoting cancer but boosting antitumor immunotherapy. J. Biomed. Sci. 2022, 29, 12. [Google Scholar] [CrossRef] [PubMed]
  185. Bojkova, B.; Winklewski, P.J.; Wszedybyl-Winklewska, M. Dietary Fat and Cancer-Which Is Good, Which Is Bad, and the Body of Evidence. Int. J. Mol. Sci. 2020, 21, 4114. [Google Scholar] [CrossRef] [PubMed]
  186. Bernard, J.N.; Chinnaiyan, V.; Andl, T.; Le Bras, G.F.; Qureshi, M.N.; Altomare, D.A.; Andl, C.D. Augmented CPT1A Expression Is Associated with Proliferation and Colony Formation during Barrett’s Tumorigenesis. Int. J. Mol. Sci. 2022, 23, 1745. [Google Scholar] [CrossRef] [PubMed]
  187. Tang, M.; Dong, X.; Xiao, L.; Tan, Z.; Luo, X.; Yang, L.; Li, W.; Shi, F.; Li, Y.; Zhao, L.; et al. CPT1A-mediated fatty acid oxidation promotes cell proliferation via nucleoside metabolism in nasopharyngeal carcinoma. Cell Death Dis. 2022, 13, 331. [Google Scholar] [CrossRef]
  188. Moon, H.S.; Batirel, S.; Mantzoros, C.S. Alpha linolenic acid and oleic acid additively down-regulate malignant potential and positively cross-regulate AMPK/S6 axis in OE19 and OE33 esophageal cancer cells. Metabolism 2014, 63, 1447–1454. [Google Scholar] [CrossRef]
  189. Pan, J.; Dai, Q.; Zhang, T.; Li, C. Palmitate acid promotes gastric cancer metastasis via FABP5/SP1/UCA1 pathway. Cancer Cell Int. 2019, 19, 69. [Google Scholar] [CrossRef]
  190. Xiang, F.; Wu, K.; Liu, Y.; Shi, L.; Wang, D.; Li, G.; Tao, K.; Wang, G. Omental adipocytes enhance the invasiveness of gastric cancer cells by oleic acid-induced activation of the PI3K-Akt signaling pathway. Int. J. Biochem. Cell Biol. 2017, 84, 14–21. [Google Scholar] [CrossRef]
  191. Zeng, X.; Zhu, M.; Liu, X.; Chen, X.; Yuan, Y.; Li, L.; Liu, J.; Lu, Y.; Cheng, J.; Chen, Y. Correction to: Oleic acid ameliorates palmitic acid induced hepatocellular lipotoxicity by inhibition of ER stress and pyroptosis. Nutr. Metab. 2020, 17, 18. [Google Scholar] [CrossRef] [PubMed]
  192. Fatima, S.; Hu, X.; Huang, C.; Zhang, W.; Cai, J.; Huang, M.; Gong, R.H.; Chen, M.; Ho, A.H.M.; Su, T.; et al. High-fat diet feeding and palmitic acid increase CRC growth in beta2AR-dependent manner. Cell Death Dis. 2019, 10, 711. [Google Scholar] [CrossRef] [PubMed]
  193. Beyaz, S.; Mana, M.D.; Roper, J.; Kedrin, D.; Saadatpour, A.; Hong, S.J.; Bauer-Rowe, K.E.; Xifaras, M.E.; Akkad, A.; Arias, E.; et al. High-fat diet enhances stemness and tumorigenicity of intestinal progenitors. Nature 2016, 531, 53–58. [Google Scholar] [CrossRef] [PubMed]
  194. Pascual, G.; Avgustinova, A.; Mejetta, S.; Martin, M.; Castellanos, A.; Attolini, C.S.; Berenguer, A.; Prats, N.; Toll, A.; Hueto, J.A.; et al. Targeting metastasis-initiating cells through the fatty acid receptor CD36. Nature 2017, 541, 41–45. [Google Scholar] [CrossRef]
Figure 1. Obesity alters the immunometabolic landscape to support gastrointestinal cancer. Obese adipose tissue secretes metabolites that have been proven to show oncogenic potential. These include insulin, insulin-like growth factor (IGF-1), glucose, cholesterol, reactive oxygen species (ROS), and pro-inflammatory cytokines. These metabolites and cytokines then enter the circulation to neighbouring gastrointestinal organs where they drive normal cells to cancer cells and ultimately to metastatic cancer cells. IL, interleukin; MAPK, mitogen-activated protein kinase; mTOR, mechanistic target of rapamycin; PI3K, phosphoinositide-3-kinase; RAS, rat sarcoma; TNFα, tumour necrosis factor alpha. This figure was created using Biorender.com (accessed on 7 August 2023).
Figure 1. Obesity alters the immunometabolic landscape to support gastrointestinal cancer. Obese adipose tissue secretes metabolites that have been proven to show oncogenic potential. These include insulin, insulin-like growth factor (IGF-1), glucose, cholesterol, reactive oxygen species (ROS), and pro-inflammatory cytokines. These metabolites and cytokines then enter the circulation to neighbouring gastrointestinal organs where they drive normal cells to cancer cells and ultimately to metastatic cancer cells. IL, interleukin; MAPK, mitogen-activated protein kinase; mTOR, mechanistic target of rapamycin; PI3K, phosphoinositide-3-kinase; RAS, rat sarcoma; TNFα, tumour necrosis factor alpha. This figure was created using Biorender.com (accessed on 7 August 2023).
Metabolites 14 00042 g001
Figure 2. Fatty acids affect adipose distribution morphology and immune cell behaviour. Saturated and monounsaturated fatty acids can dictate adipose tissue distribution and adipocyte size (hypertrophy versus hyperplastic) [41]. Additionally, fatty acids can differentially regulate immune cell behaviour including macrophages, neutrophils, B cells, T cells, and dendritic cells. ATP, adenosine triphosphate; ER, endoplasmic reticulum; ERK, extracellular signal-regulated kinase; IFN-γ, interferon-gamma; IL, interleukin; IRS-1, insulin receptor substrate 1; JNK, c-Jun N-terminal kinase; MCP-1, monocyte chemoattractant protein-1, MKK, mitogen-activated protein kinase kinase; mTOR, mechanistic target of rapamycin; NF-κB, nuclear factor kappa B; NLRP3, NOD-, LRR- and pyrin domain-containing protein 3; NK, natural killer; NO, nitric oxide; PGC-1α, peroxisome proliferator-activated receptor gamma coactivator 1-alpha; PI3K, phosphoinositide-3-kinase; PPARγ, peroxisome proliferator-activated receptor gamma; ROS, reactive oxygen species; Th, helper T cell; TLR, toll-life receptor; TNFα, tumour necrosis factor alpha; Treg, regulatory T cell; ↑, increase; ↓, decrease. This figure was created using Biorender.com (accessed on 7 August 2023).
Figure 2. Fatty acids affect adipose distribution morphology and immune cell behaviour. Saturated and monounsaturated fatty acids can dictate adipose tissue distribution and adipocyte size (hypertrophy versus hyperplastic) [41]. Additionally, fatty acids can differentially regulate immune cell behaviour including macrophages, neutrophils, B cells, T cells, and dendritic cells. ATP, adenosine triphosphate; ER, endoplasmic reticulum; ERK, extracellular signal-regulated kinase; IFN-γ, interferon-gamma; IL, interleukin; IRS-1, insulin receptor substrate 1; JNK, c-Jun N-terminal kinase; MCP-1, monocyte chemoattractant protein-1, MKK, mitogen-activated protein kinase kinase; mTOR, mechanistic target of rapamycin; NF-κB, nuclear factor kappa B; NLRP3, NOD-, LRR- and pyrin domain-containing protein 3; NK, natural killer; NO, nitric oxide; PGC-1α, peroxisome proliferator-activated receptor gamma coactivator 1-alpha; PI3K, phosphoinositide-3-kinase; PPARγ, peroxisome proliferator-activated receptor gamma; ROS, reactive oxygen species; Th, helper T cell; TLR, toll-life receptor; TNFα, tumour necrosis factor alpha; Treg, regulatory T cell; ↑, increase; ↓, decrease. This figure was created using Biorender.com (accessed on 7 August 2023).
Metabolites 14 00042 g002
Figure 3. Adipose structure and function drive tumorigenesis. Overconsumption results in lean adipose tissue expansion to either hypertrophic metabolically unhealthy adipose tissue or hyperplastic metabolically healthy adipose tissue. The metabolically unhealthy adipose creates a pro-tumorigenic environment with increased immune cell infiltration, pro-inflammatory cytokine secretion, and increased free fatty acid release, which drive progression from a benign epithelium toward tumour growth and ultimately metastasis to a greater extent than metabolically healthy adipose. EMT, epithelial–mesenchymal transition; Th, helper T cells; NK, natural killer; Treg, regulatory T cells; ↑, increase; ↓, decrease. This figure was created using Biorender.com (accessed on 7 August 2023).
Figure 3. Adipose structure and function drive tumorigenesis. Overconsumption results in lean adipose tissue expansion to either hypertrophic metabolically unhealthy adipose tissue or hyperplastic metabolically healthy adipose tissue. The metabolically unhealthy adipose creates a pro-tumorigenic environment with increased immune cell infiltration, pro-inflammatory cytokine secretion, and increased free fatty acid release, which drive progression from a benign epithelium toward tumour growth and ultimately metastasis to a greater extent than metabolically healthy adipose. EMT, epithelial–mesenchymal transition; Th, helper T cells; NK, natural killer; Treg, regulatory T cells; ↑, increase; ↓, decrease. This figure was created using Biorender.com (accessed on 7 August 2023).
Metabolites 14 00042 g003
Figure 4. Obese adipose and fatty acid effects on gastrointestinal cancers. Obese adipose increases immune cell infiltration, which then creates a tumorigenic environment on neighbouring gastrointestinal organs including the oesophagus, stomach, liver, and colon. Effects seen are an increase in inflammation, angiogenesis, proliferation, and cell differentiation, which can differ based on the specific organ. Palmitate and oleate have differential effects on oesophageal, gastric, liver, and colorectal cancer. Generally, saturated fatty acid palmitate drives cellular behaviours which may increase a tumorigenic environment to a greater extent than monounsaturated fatty acid oleate. CPT1A, carnitine palmitoyltransferase 1A; CXCL, chemokine (C-X-C motif) ligand; FABP1, fatty-acid binding protein 1; IL, interleukin; MCP-1, monocyte chemoattractant protein-1, MDC, macrophage-derived chemokine; TNFα, tumour necrosis factor alpha; VEGF, vascular endothelial growth factor; ↑, increase; ↓, decrease. This figure was created using Biorender.com (accessed on 7 August 2023).
Figure 4. Obese adipose and fatty acid effects on gastrointestinal cancers. Obese adipose increases immune cell infiltration, which then creates a tumorigenic environment on neighbouring gastrointestinal organs including the oesophagus, stomach, liver, and colon. Effects seen are an increase in inflammation, angiogenesis, proliferation, and cell differentiation, which can differ based on the specific organ. Palmitate and oleate have differential effects on oesophageal, gastric, liver, and colorectal cancer. Generally, saturated fatty acid palmitate drives cellular behaviours which may increase a tumorigenic environment to a greater extent than monounsaturated fatty acid oleate. CPT1A, carnitine palmitoyltransferase 1A; CXCL, chemokine (C-X-C motif) ligand; FABP1, fatty-acid binding protein 1; IL, interleukin; MCP-1, monocyte chemoattractant protein-1, MDC, macrophage-derived chemokine; TNFα, tumour necrosis factor alpha; VEGF, vascular endothelial growth factor; ↑, increase; ↓, decrease. This figure was created using Biorender.com (accessed on 7 August 2023).
Metabolites 14 00042 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mitchelson, K.A.J.; O’Connell, F.; O’Sullivan, J.; Roche, H.M. Obesity, Dietary Fats, and Gastrointestinal Cancer Risk-Potential Mechanisms Relating to Lipid Metabolism and Inflammation. Metabolites 2024, 14, 42. https://0-doi-org.brum.beds.ac.uk/10.3390/metabo14010042

AMA Style

Mitchelson KAJ, O’Connell F, O’Sullivan J, Roche HM. Obesity, Dietary Fats, and Gastrointestinal Cancer Risk-Potential Mechanisms Relating to Lipid Metabolism and Inflammation. Metabolites. 2024; 14(1):42. https://0-doi-org.brum.beds.ac.uk/10.3390/metabo14010042

Chicago/Turabian Style

Mitchelson, Kathleen A. J., Fiona O’Connell, Jacintha O’Sullivan, and Helen M. Roche. 2024. "Obesity, Dietary Fats, and Gastrointestinal Cancer Risk-Potential Mechanisms Relating to Lipid Metabolism and Inflammation" Metabolites 14, no. 1: 42. https://0-doi-org.brum.beds.ac.uk/10.3390/metabo14010042

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