v
Search
Advanced Search

Publications > Journals > Gene Expression > Article Full Text

  • OPEN ACCESS

Fructose, Another Sweet for Cancer: A Context Acting Nutrient Hypothesis

  • Tomas Koltai1 and
  • Larry Fliegel2,* 
 Author information
Gene Expression   2023;22(2):141-155

doi: 10.14218/GE.2023.00014

Abstract

Rapidly proliferating cancer cells exhibit a high energy demand. However, their utilization of the glycolytic pathway is inefficient, leading to a compensatory effect wherein cancer cells consume ten to twenty times more glucose than normal cells. In cases where glucose availability is limited due to a poorly perfused hypoxic microenvironment, cancer cells resort to alternative energy sources, including fructose. Certain tumors have been found to rely heavily on fructose, and fructose utilization contributes to pro-tumoral signaling and increased cancer risk. Over the past 70 years, dietary fructose intake has steadily increased, resulting in a rise in obesity and metabolic syndrome, both of which elevate cancer risk. In this paper, we present compelling evidence that highlights the role of fructose and the glucose transporter GLUT5 in promoting specific types of tumors. We summarize the existing evidence and pathways through which fructose contributes to cancer metabolism, particularly in cases where glucose availability is restricted. Furthermore, we propose a hypothesis that elucidates the regulation of the lipogenic phenotype by dietary fructose intake and cellular energy status. It is important to note that the effects of fructose are context-dependent, with its tumor-promoting effects varying based on the energy status of the cell. We comprehensively analyze why targeting fructose uptake and fructolysis should be important for the management of some tumors and cancer prevention.

Keywords

Cancer, Fructose, Glucose, GLUT5, Tumors

Introduction

Glucose is classically considered the primary source of energy for cancer cells.1–5 Glucose uptake by malignant cells is more than ten times higher than in normal cells.6–9 The increased tumor glucose uptake has led to the development of a very useful diagnostic method to visualize them: the 18F-2-fluoro-2-deoxyglucose positron emission tomography (PET).10

Paradoxically, despite the much higher glucose uptake in malignant cells, tumor glucose concentrations are much lower than in their normal counterparts.11,12 However, this apparent paradox has an explanation. There is no paradox: all the glucose taken up in tumors is swiftly used by the elevated metabolism of tumor cells, and thus very little remains in the tumor microenvironment. PET, mentioned above, uses the derivative of glucose, 18F-2-fluoro-2-deoxyglucose, to detect tumor cells. The reason why PET shows high 18F-2-fluoro-2-deoxyglucose concentrations in tumors is that this derivative of 2-deoxyglucose blocks glycolysis, and there is no further degradation,13 allowing accumulation of an 18F-2-fluoro-2-deoxyglucose for visualization.14

The increased glucose uptake of tumors, and the fact that the arrival of nutrients to a very hypoxic, poorly perfused tumor is limited, should prompt a question: how can cancer cells get all the necessary energy needed from glucose in a depleted environment that is difficult to replenish? Beyond glutamine as a proxy, this question has not yet been fully answered. One possibility is that fructose is an important energy source in at least some types of cancer, and there is evidence to support this concept.15–18 However, evidence suggests that fructose uptake is significantly reduced in certain tumors, such as hepatocarcinoma, compared to normal liver tissue.19 So perhaps, there is a situation whereby in most tumors, fructose is an important contributor to tumor growth and metabolism, but this is not the case in hepatomas.

Interestingly, the situation may vary in different types of tumors. In prostate cancer, glucose is not the primary energy source, at least in the early stages.20 This explains why PET studies based on glucose uptake in prostate cancer have a low diagnostic value.21–23 Fructose is transported passively across cell membranes by the glucose transporter GLUT5,24 whereas GLUT1 is selective for glucose.25 It has been shown that prostate high-grade intraepithelial neoplasias show increased expression of GLUT5 and an absence of GLUT1, suggesting that early prostate malignancies use fructose rather than glucose as an energy source.26,27 It seems that the importance and uptake of fructose may vary from one type of malignancy to another.

Fructose metabolism

To understand the role of fructose in tumor malignancy, it is crucial to understand its metabolism compared to glucose. Fructose and glucose are metabolized differently. Both require transporters to cross the cell membrane, but glucose is taken up by glucose transporters 1 to 4 (GLUT 1–4), while fructose requires the specific transporter GLUT5.28,29 The tissue uptake is also different for the two sugars. Almost all fructose is normally metabolized in the liver (80%), while glucose can be metabolized in any tissue.30 Another difference is in the first step of metabolism. Although the fructose molecule is very similar to glucose, it is not readily phosphorylated by hexokinases that phosphorylate glucose.

Cells that can metabolize fructose, therefore, use a ketohexokinase called fructokinase. This is followed by a step with aldolase B (Fig. 1).31–34 The two steps in fructolysis begin with fructose phosphorylation by the enzyme fructokinase. ATP is used to donate the phosphate group producing fructose-1-phosphate. In the second step, another enzyme, aldolase B splits the molecule into two trioses: dihydroxyacetone phosphate and glyceraldehyde. The other parts of the pathway are like glucose metabolism. Glyceraldehyde needs to be phosphorylated to continue its metabolism, and this is done by a triokinase-producing glyceraldehyde 3-phosphate. For glycolysis or oxidative metabolism pathways, dihydroxyacetone phosphate is changed to glyceraldehyde 3-phosphate by an isomerase, and now the two glyceraldehyde 3-phosphate trioses can follow the same steps (Fig. 1).

Fructose metabolism.
Fig. 1  Fructose metabolism.

Fructose is transported into the cell by GLUT5 and is phosphorylated by Fructokinase, also known as ketohexokinase. Aldolase B splits the molecule into two trioses: dihydroxyacetone phosphate and glyceraldehyde. Glyceraldehyde is phosphorylated by a triokinase-producing glyceraldehyde 3-phosphate. Dihydroxyacetone phosphate is changed to glyceraldehyde 3-phosphate by an isomerase. “1” represents oxidative metabolism, and “2” the glycolytic pathway. After the glyceraldehyde 3-phosphate stage is reached, glucose and fructose metabolism merge. However, fructose achieves this stage without passing through the phosphofructokinase enzymatic action as glucose metabolism does. Phosphofructokinase (PFK) is a metabolic checkpoint under insulin control. ATP, adenosine triphosphate; ADP adenosine diphosphate; PFK, phosphofructokinase.

Figure 1 also shows the pathway leading to lipid synthesis from fructose. The pathway that leads to glycogen formation in the liver is not shown. The fact that fructose is mainly metabolized in the liver and normally only to a minor extent in other tissues means that when tumors use fructose as an energy source, these malignant tissues develop normal hepatic abilities.

There are notable differences in regulation between glycolytic enzymes and fructolytic ones. Firstly, the hexokinase enzymes (except for the hexokinase IV, or glucokinase, expressed in the liver and kidney tubular cells, enterocytes, and pancreatic alpha- and beta-cells) are inhibited by increased concentrations of their product, glucose-6-phosphate. Second, phosphofructokinase (PFK) is tightly regulated in glucose metabolism, and fructose, arrives at the triose pool bypassing PFK without these restrictions.31,35 Therefore, trioses produced from fructolysis can swiftly and unrestrictedly generate a substrate for the pathways shown in Figure 1, including lipogenesis.

Another metabolic problem with fructolysis is caused by the rapid phosphorylation of fructose to fructose-1- phosphate. If it is not matched by downstream ATP production from fructose-1-phosphate catabolism, this can lead to a drop in hepatic ATP stores and a rise in hepatic inorganic phosphorus concentration. This can cause acute hepatic dysfunctions due to hepatocytes’ energy deprivation, resulting in hypoglycemia and increased uric acid production. Additionally, fructose-1-phosphate can indirectly stimulate glucokinase by activation of a glucokinase-regulating protein. This effect may contribute to hypoglycemia development by decreasing hepatic glucose production (Fig 2).35,36

Increased uric acid production due to increased fructose catabolism.
Fig. 2  Increased uric acid production due to increased fructose catabolism.

AMP, adenosine monophosphate; ATP, adenosine triphosphate; IMP, inositol monophosphate.

There are two possible pathways for fructose (Fig. 1) which leads us to a question: what makes uncontrolled fructolysis go one way or the other? A possible answer is that fructolysis goes toward energy production when energy requirements are not met. Otherwise, it goes toward lipogenesis. This needs experimental confirmation and is further discussed in section “I” below.

Fructose-1-phosphate is an allosteric activator of pyruvate kinase, representing glycolysis’ last step.37 It also activates transcription factors such as Sterol regulatory element binding protein-1 (SREBP1) and Carbohydrate Responsive Element Binding Protein(ChREBP). SREBP-1 is a transcription factor for genes that participate in glucose metabolism and lipogenesis.38 It is involved in the growth and progression of prostate cancer,39 promotes migration and invasion of breast cancer,40 regulates fatty acid synthase,41 which is also pro-carcinogenic, promotes invasion and metastasis in hepatocarcinoma,42 among other pro-tumoral effects. Activation of SREBP-1 by fructose-1-phosphate may promote tumors as it has been shown that down-regulation/inhibition of SREBP-1 has antitumoral effects in glioblastoma43 and other tumors.44 The other transcription factor activated by fructose-1-phosphate, ChREBP, stimulates many glycolytic and lipogenic enzymes that are potentially important in cancer progression.45 It contributes to cell proliferation46 and aerobic glycolysis.47,48 It also regulates the androgen receptor transcription in prostate cancer.49

Fructose is obtained from the diet, and how dietary input is handled is quite interesting. Short-term elevations in the human diet of fructose increase hepatic glucose production. They also increase basal and postprandial blood triglyceride concentrations and intrahepatic fat content. These metabolic alterations may be early markers of metabolic dysfunction or adaptations to the specific two-step fructose metabolism.35 Dietary fructose is efficiently absorbed in the lower duodenum and jejunum and is processed in the liver. GLUT5 serves as the primary transporter responsible for fructose absorption. In mice, the deletion of GLUT5 leads to a significant reduction of 75% in fructose absorption in the jejunum, as well as a substantial decrease of 90% in serum fructose levels.50 Absorbed fructose circulates in the serum and is delivered to the liver and to other tissues. Some tumors over-expressGLUT551–53 and this is an indirect sign that these tumors have developed the ability to absorb fructose from the serum and are obtaining part of their energy from fructose.

Evidence of fructose as an important source of energy in cancer cells

There is abundant evidence that fructose can be an essential energy source for cancer cells. We divide this evidence into six general types. Firstly, many studies have demonstrated that the fructose transporter GLUT5 is upregulated in cancer cells. This was shown in breast cancer cells,18,54 where it was also demonstrated that GLUT5 is almost absent in normal cells.54 Other cancers in which GLUT5 was upregulated include clear cell renal carcinoma,53 ovarian cancer tissues,52 acute myeloid leukemia cells,51 microglia of human gliomas,55 lung tumor tissue of patients with adenocarcinoma,56 human colorectal cancer specimens,57 Philadelphia positive acute lymphoblastic cells,58 endothelial cells from hepatocellular carcinoma,59 glioma cells60 and colorectal cancer cells.61 The exact mechanism by which GLUT5 expression is increased is not completely clear. There are some reports of factors that may be involved. Medina Villaamilet et al.62 found a correlation between HIF-1α (hypoxia-inducible factor 1-alpha, a transcription factor that regulates angiogenesis and tumor growth and metastasis) and GLUT5 expression. Another study showed that the inflammatory IL-6/STAT3 axis activates GLUT5 regulating the fructose metabolism in oral squamous cell carcinoma cells and prostate cancer cells. STAT3 transcription factor binds the GLUT5 gene’s promoter region, enhancing its transcription.63 However, it is unclear if these are the only factors involved, and details of GLUT5 induction pathways remain to be elucidated. However, several studies have demonstrated that increased GLUT5 expression results in increased fructose uptake. This was shown in acute myeloid leukemia cells,51 clear renal cell carcinomas53 and glioma cells.60 In several cases the increase in fructose uptake was correlated with factors such as malignant progression and differentiation.53 To summarize, many studies in different cancer types have shown that GLUT5 expression is elevated, and several other studies have confirmed these results in increased fructose uptake.

The second type of evidence supporting a fructose role in tumors are several studies showing that knockdown or inhibition of GLUT5 has inhibitory effects on tumors’ cell growth, viability, migration, and proliferation. These studies were demonstrated in various cancer types and different circumstances. An example of a competitive inhibitor of fructose transport by GLUT5 is 2,5-anhydro-D-mannitol, and administration of this compound could markedly suppress clear cell renal cell carcinoma growth.64 Similarly, in acute myeloid leukemic cells, pharmacological blockade of fructose uptake with the same compound weakened the malignant phenotype and increased cell sensitivity to chemotherapeutic drugs,51 while in colon cancer cells, a different GLUT5 inhibitor N-.4-(methylsulfonyl)-2-nitrophenyl-1,3-benzodioxol-5-amine, significantly decreased viability of these cancer cells but had little effect on the viability of normal colon epithelium cells.57 Knockdown of GLUT5 may be a more specific method of reducing GLUT5 activity, and several studies have shown that this has beneficial effects on preventing tumor cell growth and malignancy traits. For example, Jin et al.64 deleted the GLUT5 gene from clear cell renal cell carcinomas cells. Cell malignancy was attenuated, and apoptosis was activated. Similar results were shown with the knockdown of GLUT5 in glioma cells,60 ovarian cancer cells,52 and two types of human breast cancer cells, MCF-7 and MDA-MB-231 cells.65 Very recently, Groenendyk et al.66 also demonstrated that CRISPR/Cas9 mediated inactivation of the SLC2A5 (GLUT5) gene inhibited cancer cell proliferation and migration in vitro, as well as metastases in vivo in several different animal models. Moreover, SLC2A5 attenuated cells significantly altered mitochondrial architecture and localization, indicating an important role in directing mitochondrial function for cancer cell motility and migration. The study used MIA PaCa-2 cells, a highly metastatic pancreatic ductal adenocarcinoma cell line and HT-1080 human cells derived from connective tissue of a patient with fibrosarcoma. Overall, these types of studies demonstrate that inhibition or knockdown of GLUT5 affects several different types of cancer cells, having beneficial effects such as inhibiting tumor growth and survival. These studies confirm that induction of expression of the GLUT5 gene in these cancers, has real and significant effects with regards to tumor progression.

A third line of evidence that supports the vital role of fructose as an energy source for tumors are the effects of high fructose in the “diet”of cells, animals, or humans. Several results show that in animal models, elevated dietary fructose enhances carcinogenesis. For example, a dietary treatment with fructose increased hepatocarcinogenesis in a rat model treated with N-nitrosomorpholine.67 The same group used a similar model to show that high dietary fructose enhanced nodules of atypical acinar nodule cells, which are precursors of pancreatic lesions.68 Elevated levels of fructose in the “diet” of cells could support breast cancer cell proliferation when glucose levels were reduced.18 In a separate breast cancer model, substituting the energy source in MDA-MB-468 breast cancer cells with fructose induced a more aggressive phenotype characterized by enhanced migration and invasion capabilities.16 Effects of high fructose in the medium of cells were shown in several other studies, including glioma cells, where it promoted tumor progression and GLUT5 expression.60 Another paper69 examined many different cell types and showed that cells chronically cultured in fructose develop high fructose lysis ability. The SLC2A5 (GLUT5 transporter) gene was specifically upregulated, as was fructose usage. Fructose elevated GLUT5 expression and stimulated cell proliferation.

Elevated levels of fructose in animal diets have also been shown to enhance tumorigenesis in several animal models. This occurred in lung metastasis and mammary gland tumorigenesis70 and a mouse model of hepatocarcinoma.71 Kuehm et al.72 also demonstrated that in melanoma tumors in the C57BL/6 mouse model of diet-induced obesity, dietary fructose promoted cytoprotection and resistance to immunotherapy. Mice with a high fructose diet had increased expression of the cytoprotective enzyme heme oxygenase-1, which shielded tumor cells from immune-mediated killing. The increase of this protein was recapitulated in human A375 melanoma cells exposed to fructose in culture. Another recent publication also showed that dietary fructose improves the survival of intestinal cells and increases villus length in several mouse models.73

The elevation of fructose in the diet of humans has been analyzed, mainly by examining the incidence of various tumors in populations or groups with high fructose input. For example, a study of over 80,000 women74 found an association between higher fructose intake in obese, sedentary women with increased pancreatic cancer risk. Another association was published by Larsson et al.,75 who found increased pancreatic cancer risk in obese women with high consumption of sweetened soft drinks (sweetened soft drinks contain a high level of fructose). Moreover, a different sizeable multiethnic cohort study showed an increased relative risk for pancreatic cancer in people with high fructose intake.76 Finally, a retrospective analysis of a large population, including ten cohorts, did not show an association between pancreatic cancer risk and intake of diets high in glycemic index, glycemic load, total carbohydrates, or sucrose. However, there was enhanced pancreatic cancer risk with high fructose diets.77 These studies support the concept that aside from all the animal studies and studies in culture described above, humans are also susceptible to elevated levels of fructose in their diet.

Here we also suggest that the levels of fructose updated in the “diet” can compensate for decreased or blocked glucose usage and stimulate glucose usage. Cancer cells have significantly elevated levels of glycolysis and high metabolic activity with subsequent high energy usage.78 As glucose is the often-used substrate of cancer cells, one would believe that blockage of glycolysis should be a useful anti-cancer strategy. 2-deoxyglucose blocks glucose metabolism; however, it has failed in tests as an anti-cancer drug.79 We suggest that the failure of anti-glycolytic approaches may be due to using fructose as an alternative energy source. Fructose metabolism bypasses the glycolytic obstruction induced by 2-deoxyglucose. One example was a study demonstrating that fructose was an important fuel for lung adenocarcinoma when glucose levels were low.80 However, it is also interesting that fructose metabolism can stimulate the glycolytic pathway, which can stimulate cancer growth and metastasis.15 Thus, fructose can substitute for glucose where needed and stimulate glucose use if available, thus promoting tumor progression.

A fourth line of evidence that supports the critical role of fructose in human cancer development is the clinical correlation that suggests that fructose metabolism is important in human tumor progression. For example, GLUT5 expression is elevated in human glioma tissues, and GLUT5 is correlated with glioma progression and poor survival in glioma patients.60 A similar correlation of GLUT5 levels with ovarian cancer tumor malignancy and progression-free survival has also been reported,52 GLUT5 was also upregulated in lung adenocarcinoma patients and correlated with poor prognosis.56 There is an almost 2.5-fold (p < 0.001) increase in GLUT5 mRNA expression level in colorectal cancer specimens compared with the healthy intestinal mucosa.57 GLUT5 was also overexpressed in a survey of 215 different human tumor samples.81 Immunolocalization studies revealed that GLUT5 is highly expressed in vivo in human breast cancer but is absent in normal human breast tissue.54 Similar to the effect with GLUT5, ketohexokinase, a key enzyme of fructose catabolism, is over-expressed in gliomas. This overexpression correlated with tumor progression and poor survival of glioma patients.82 Thus, in several different types of cancers, GLUT5 and another key enzyme of fructose metabolism, ketohexokinase, are upregulated and this elevation correlates with detrimental effects on human health. This indicates that fructose metabolism is linked to cancer progression in humans. On its own, clinical correlates of expression with cancer progression are only indirect evidence at best, but when one considers the other data above such as knockdown and inhibitor studies in animals and cells described above, this provides good evidence for an important role of fructose metabolism in tumor progression.

A fifth line of evidence supporting fructose’s role in cancer progression are the metabolic alterations or reprogramming that occur in many tumors that allow them to use fructose more efficiently. For example, colon cancer cells overexpress aldolase B when metastasizing in the liver. Aldolase B enhances fructose metabolism. Restricting dietary fructose or targeting aldolase B decreased the growth of liver metastases without affecting the primary tumor.83 Another enzyme that is altered is transketolase. Fructose can induce transketolase flux which promotes pancreatic cancer.84 Through increased fructose in dietary sugar, the lipoxygenase pathway in mice was increased, raising the risk of breast cancer development and metastases.70 In another study, it was shown that fructose, but not glucose, reprogrammed malignant human prostate cancer cells. It significantly altered mRNA expression of Hexokinase 2, type-C fructokinase, pyruvate kinase M2 and type-A lactate dehydrogenase. Various metabolic alterations improving fructose utilization also occurred in pancreatic stem cells,85 and prostate cancer cells using fructose as a main energy source when GLUT5 was overexpressed.86 In tumor endothelial cells from hepatocellular carcinoma, fructose treatment promoted proliferation, migration and angiogenesis. Fructose metabolism was elevated and both GLUT 5 and ketohexokinase were upregulated. Knockdown or inhibition of these proteins abolished fructose-induced tumor angiogenesis and suppressed tumor growth.59 Demonstration of elevation of ketohexokinase in response to elevated dietary fructose has been shown several times. This also occurred in fructose fed mice87 and in colon cancer cells, where GLUT5 expression inhibited ketohexokinase degradation.61 In hepatocellular carcinoma, fructose promoted aggressiveness, and mice with a fructose enriched diet had appropriate metabolic reprogramming that increased energy, NADPH, and nucleotide production, allowing for increased tumor cell aggressiveness.88

A sixth line of evidence shows the relationship between fructose metabolism and tumor immunology. Kuehm et al.72 found that mice melanoma tumors in animals with a high-fructose diet were resistant to immunotherapy. They also found increased expression of heme oxygenase-1, a cytoprotective enzyme to which they attributed possible participation in the process. Interestingly, when exposed to fructose, A375 melanoma cells in culture showed high heme oxygenase-1 expression. This expression was causally linked to resistance to immune checkpoint inhibitors.

Based on the above-mentioned reports, there is evidence of a fructose-dependent pathway and a GLUT5-dependent pathway in cancer that stimulate tumorigenesis, as shown in Figure 3.

Relationship among fructose metabolism, glut5, hexokinase and tumorigenesis.
Fig. 3  Relationship among fructose metabolism, glut5, hexokinase and tumorigenesis.

The above information makes it clear that fructose uptake and GLUT5, the fructose transporter, are important in at least some types of cancer and facilitate the growth and proliferation of cancer cells.

Fructose levels in the bloodstream

Notably, the peripheral plasma fructose concentration is relatively low (approximately 0.04 mM). After fructose ingestion, GLUT5 expression is increased in the intestine.89,90 Fructose levels can increase 10-fold and return to normal after 2 hours fasting.91,92 Fructose levels in the bloodstream have also been shown experimentally to increase in response to fructose ingestion. This was shown in humans with acute fructose administration93 and with the administration of high-fructose corn syrup-sweetened soft drinks, where dramatic increases in fructose concentrations were shown with ingestion.94 It is also important to note that the fasting serum concentration of fructose is significantly higher in pancreatic cancer patients than in healthy individuals.95

Further details on the relationship of GLUT5 with cancer

Localization

GLUT5 is a fructose transporter that facilitates the diffusion of fructose in a concentration-dependent manner.96 It is highly expressed on the apical border of intestinal mucosa cells. It has modest expression levels in other tissues such as adipocytes, kidneys, and skeletal muscles,97,98 while other tissues have no or minimal expression of GLUT5. As noted above, some tumors express GLUT5, while their corresponding healthy tissues do not express GLUT5. GLUT5 is the vehicle for the rapid absorption of fructose into cells, and the liver metabolizes approximately 70–80% of the absorbed fructose. Very little is known about GLUT5 regulation in tumors. It is unclear if the regulatory proteins that work on GLUT5 in some tissues (Fig. 4) are operative in tumors. As always, suspicion is cast on several putative tumor drivers, but experimental evidence is lacking (Fig. 4).24,89–105

Regulatory mechanisms of GLUT5.
Fig. 4  Regulatory mechanisms of GLUT5.

These seem to work in the intestine and other tissues. a, dietary fructose elevates GLUT5 expression;89,99,100 b, insulin treatment can increase GLUT5 expression including elevated transcription and protein levels;101 c, Developmental reprogramming and induction of rat GLUT5 requires glucocorticoid receptor translocation to the nucleus;24 d, carbohydrate responsive element-binding protein is a transcription factor that regulates GLUT5 expression in response to carbohydrates;102 e, fully differentiated Caco-2/TC7 (human colon adenocarcinoma) cells, thyroid hormone, and glucose increase GLUT5 mRNA abundance in dose-dependent manners;103 f, In human adipocytes hypoxia increases GLUT5 expression markedly.104 Some of these mechanisms may work in tumors such as hypoxia,105 but further studies are required to better characterize this in tumors.

Regulation of expression

Tumors are highly hypoxic, and this may be the cause of GLUT5 over-expression. Hypoxia in adipocytes was demonstrated to increase GLUT5 levels.104 Additionally, hypoxia elevated levels of GLUT5 mRNA and protein in breast cancer cells. However, this depended on the type of breast cancer cell.105 Breast cancer samples from patients showed increased expression of all GLUTs and HIF-1α compared to control tissue. Triple-negative breast cancer tends to be more severe than other types. In triple-negative breast cancer, the samples showed greater GLUT1, GLUT5, and HIF-1α expression levels than ER-positive cases.105

Godoy et al.81 studied GLUT isoform expression in different benign and malignant tumors. GLUT1 was the main isoform detected in tumor tissues. However, GLUT5 was extensively expressed in malignant tumors suggesting that fructose transport for use as an energy substrate was widespread. GLUT5 was detected in colon adenocarcinoma, ependymomas, plexus choroids papilloma, lung mesothelioma, liver carcinoma, lymphomas (only GLUT5), testis seminoma, and uterus leiomyoma. GLUT5 was also detected in breast tumor cells and was more intense than normal cells, especially in invasive ductal carcinoma samples. GLUT2, a lower affinity fructose transporter, was similarly expressed in elevated amounts in tumor tissues when GLUT5 expression was elevated.

Downstream effects of GLUT5 expression

GLUT5 also plays an important role in lung cancer cell growth where the tumor seems to be fructose dependent. Chen et al.106 found that by up-regulating GLUT5, these cells could use fructose as an alternative to glucose in vivo. Importantly, this fructose was mainly used for lipid synthesis. Deletion of the GLUT5 gene (SLC2A5) impaired proliferation, which could be restored by administering fatty acids. They also showed that GLUT5-mediated fructose import was necessary to inhibit AMPK, thus allowing mTORC1 to promote lung cancer growth.

GLUT5 expression and chemoresistance

Several studies have shown that elevated expression of GLUT5 induces resistance to chemotherapy or correlates with beneficial effects (from the tumor’s point of view). For example, Ramzy et al.107 demonstrated that GLUT5 was significantly upregulated in colorectal cancer cells and this induced drug resistance to chemotherapy treatments. A similar effect on chemotherapy resistance was also found by Shen et al. in colorectal cancer cells.61 A different study on colorectal cancer cells108 suggested that regulation by AKT was responsible for elevated GLUT5 expression and resistance to chemotherapy. Silencing either AKT or GLUT5 expression attenuated migration, invasive behavior, and inhibition of GLUT5 activity with 2,5-anhydro-d-mannitol re-sensitized these cells to chemotherapeutic treatments. These results were more than a correlation of behavior and, more specifically, confirmed the role of GLUT5. It is worth mentioning that another study109 also showed a connection between AKT, GLUT5, and cancer cell migration. In that study in lung cancer cells, overexpression of GLUT5 promoted cell migration and AKT activation. Again, the use of GLUT5 inhibitors blocked cell migration and AKT activation. There is clearly a link between GLUT5 and AKT and cancer cell behavior, including migration and chemoresistance.

Ketohexokinase (hepatic fructokinase)

As noted above, ketohexokinase is elevated in several types of cancer, and reduction or inhibition of the protein can inhibit the malignant phenotype. Here we describe some further details about this enzyme. Ketohexokinase (KHK) is the enzyme that converts fructose to fructose-1-phosphate110,111 using one molecule of ATP or GTP as a cofactor (Fig. 1). This occurs in the presence of K+ and Mg++ with no rate variations in enzyme activity within the pH range of 6 to 9. The reverse reaction occurs only at an acidic pH between 5 and 6, which is usually impossible in the cell. There are two KHK isoforms, KHK-A and KHK-C. They are generated through mutually exclusive alternative splicing of KHK pre-RNAs. KHK-C displays greater affinity for fructose compared with KHK-A, and KHK-C is produced primarily in the liver. This restricts fructose metabolism almost exclusively to this organ.112,113 Mirtschink et al.113 found that hypoxia can induce changes in splicing, switching KHK-A to KHK-C isoform in the myocardium and thus enforcing fructose metabolism. This has not been investigated in tumors, but we may speculate that something similar may happen. In addition to KHK canonical enzymatic function, it also has non-canonical activities as a protein kinase.114

Ketohexokinase acts as a nuclear kinase that has pro-tumoral effects such as:

  • Promoting tumor progression of glioma;82

  • Promoting progression of non-small lung cancer;115,116

  • Promoting pancreatic cancer growth by activating MAP Kinases pathway;117

  • Driving hepatocellular carcinoma formation by a cMyc-induced splicing switch to isoform KHK-A;118

  • Promoting fructose-induced metastasis of breast cancer.87

In summary, ketohexokinase consists of two isoforms, and the relative amount of each varies. Aside from phosphorylating fructose and beginning the fructolytic pathway, the kinase has other pro-tumoral effects contributing to cancer growth and metastasis.

Interventions to inhibit fructose as a source of energy in cancer

Given that fructose can be an important source of energy for cancer and that fructose utilization can stimulate glycolysis and may have other pro-tumoral effects, it is important to discuss how to prevent these dangerous consequences of fructose metabolism. There are several possible approaches to prevent the use of fructose as an energy source in cancer. These include:

  • Dietary restriction of fructose.

  • Drugs such as 2,5-anhydro-D-mannitol (2,5-AM), a fructose analog with a high affinity for GLUT 5 that acts as a competitive inhibitor.119

Dietary modifications for enhanced cancer therapy have been considered.120 Mostly, these have been concerned with treatments such as fasting and glucose restriction. Fasting has long been considered in mice, where combining fasting cycles with chemotherapy improves responsiveness. However, this is quite a harsh therapy. Glucose has many tumorigenic roles, as discussed above, and glucose restriction has been tested and has value. However, cells seem to be able to rewire their metabolic programs in response, as noted above.120 Fructose restriction has not been as well studied, but some evidence has suggested that, at least in mouse models, it can have beneficial effects.120,121

As noted above, compounds such as 2,5-anhydro-D-mannitol can theoretically have beneficial effects, such as re-sensitizing cells to chemotherapeutic treatments.108 The idea of targeting fructose metabolism in tumors that over-express GLUT5 is not overly complicated because fructose-restricted diets are well tolerated, and the uptake inhibitor 2-5-anhydro-D-mannitol is a non-toxic compound. However, further study is needed.

Other compounds that have been reported to have GLUT5 inhibitory abilities are:

  • MSNBA (N-[4-(methylsulfonyl)-2-nitrophenyl]-1,3-benzodioxol-5-amine);122

  • Conjugates of 2,5-anhydro-mannitol also inhibit GLUT5.

  • 6-O-allyl-d-fructofuranose;123

  • 1,3-oxazolidin-2-thiones and 1,3-oxazolidin-2-ones;123

  • Green tea and chamomile tea;124

  • allylamine derivative of 2,5-anhydro-d-mannitol;119

  • Flavonoids: epigallocatechingallate and apigenin but not quercetin;125

  • Astragalin-6-glucoside;126

  • None have been clinically tested.

When to treat

Patients with tumors expressing high levels of GLUT5 might benefit from both the above-mentioned treatment strategies. There is clear evidence of this in acute myeloid leukemia.51 In other tumors like breast, colon, and pancreatic carcinomas expressing GLUT5, treatments modulating fructose intake or uptake would probably delay growth and metastases, but this needs verification.

General Discussion

Glucose is the main energy source of cancer cells. However, tumors are often poorly perfused and the concentration of nutrients such as glucose is often lower in tumors than in normal tissues. Thus, glucose depletion affects tumors’ ability to proliferate. Different cancer cell lines show different sensitivities to glucose depletion.127 With the effects of glucose depletion, in certain tumors, fructose becomes a very important energy source and can even replace glucose as the main nutrient. Fructose consumption is associated with a more malignant phenotype with increased proliferation, invasion, and metastasis.128 However, in 2009, an epidemiological study could not confirm that high dietary fructose intake increased cancer risk.129 Nevertheless, other studies have shown some associations between elevated fructose and cancer frequency, but this is not always true.130,131 Though these results are conflicting (see also below), they do not mean that fructose has no role in cancer development. At the clinical level, fructose indirectly participates in cancer through obesity and metabolic syndrome. At the molecular level, fructose has an important role in the development of a more aggressive cancer phenotype in some tumors.18,27

The importance of fructose in cancer seems to stem from three different roles:

  • As a source of energy;

  • Reprogramming of cellular metabolism;

  • Stimulating synthesis of fatty acids necessary for the lipogenic phenotype.

Here we review some of the more critical aspects of these three capabilities of fructose.

Fructose as a source of energy

High intake of dietary fructose is increasingly being considered a causal factor of obesity,132 metabolic syndrome,133–135 and indirectly, insulin resistance.136 While studying metabolic syndrome, interesting clues were discovered on the role of fructose beyond its energetic function. Metabolic syndrome and insulin resistance increased the expression of interleukin-6,137 Akt, NF-kB,138 and TNFα via hepatic production. All these compounds have pro-tumoral activity. Thus fructose, while promoting metabolic syndrome, obesity, and insulin, simultaneously increases the production of compounds that promote tumorigenesis. These findings have not been mechanistically well characterized. Fructose supplementation in the diet has been shown to impair signaling in insulin-sensitive tissues (Fig. 4).139

Whatever the cause, metabolic syndrome and obesity can increase cancer risk. One study of 38,940 cases of cancer140 found that metabolic syndrome had an increased relative risk(RR) for hepatic (RR = 1.43), colorectal (RR = 1.25), and bladder (RR = 1.10) cancers in men, and endometrial (RR = 1.64), pancreatic (RR = 1.58), postmenopausal breast (RR = 1.56), and rectal (RR = 1.52) cancers in women. Another prospective study of 90,000 people beginning without cancer showed that in obese people, the risk of death from cancer was 52% (RR = 1.52) higher in men and 62% (RR = 1.62) higher in women, compared with the non-obese population.141

It should be noted that dietary fructose intake has increased dramatically in the United States, and many authors consider that this is partly responsible for the obesity epidemic.102,142–148 However, other authors do not accept this as the cause of energy over-consumption, independent of fructose.149,150

Fructose reprogramming cellular metabolism

Fructose can increase aerobic glycolysis (bypassing glycolytic restrictions and activating pyruvate kinase) and lipogenesis. In prostate cancer, a lipogenic phenotype was found in many tumors.151 De novo lipogenesis is a consequence of androgenic stimulation of SREBP.152 De novo lipogenesis is also one of the effects of fructose (Fig. 4). Prostate cancer patients treated with androgen deprivation can retain their lipogenic phenotype thanks to the activity of fructose metabolites as suggested by Carreño et al.153 Support for this concept came in one study that showed that adding fructose to the medium of cultured adipocytes increased lipogenesis.154 Fructose can also induce insulin resistance by activating the peroxisome proliferator-activated receptor γ coactivator-1 β (PGC-1β), which is a co-activator of SREBP-1. Knockdown of PGC-1β improved insulin resistance.155

Fructose and the lipogenic phenotype

Fructose is implicated in the lipogenic phenotype, insulin resistance, and metabolic syndrome.156 Increased fructose intake seems to be directly responsible for enhanced obesity and de novo lipogenesis.157 However, a panel of experts convened by The Center for Food, Nutrition, and Agriculture Policy was asked to examine the scientific literature on the relationship between high fructose intake and obesity. They concluded that high dietary fructose intake “does not appear to contribute to obesity any differently than do other energy sources”.158 This makes us speculate that the lipogenic phenotype induced by fructose appears only if energy requirements are fulfilled first.

Based on the above evidence, Figure 5 was constructed assuming that the energy requirements were adequately fulfilled. If this were not the case, fructose would only be used to produce energy. The cause of the switch from energy production to lipoprotein production is the energy balance. This is probably gauged by AMPK (AMP kinase). Activation of AMPK through a low ATP/AMP ratio (through the liver kinase B, LKB) inhibits ACC1 and ACC2 (Acetyl CoA Carboxylase) activity, inhibiting lipogenesis.159–161 The results of Woods et al.161 support this hypothesis. They demonstrated that chronic activation of AMPK through experimental mutation impedes lipogenesis in the liver of mice with a high fructose diet. The chronic activation of AMPK produces a surplus of ATP (Fig. 5).

Possible relationships between high fructose intake, insulin resistance, obesity, metabolic syndrome, lipogenic phenotype, and cancer.
Fig. 5  Possible relationships between high fructose intake, insulin resistance, obesity, metabolic syndrome, lipogenic phenotype, and cancer.

Fructose is a key figure that integrates these concepts. High fructose intake can lead to metabolic syndrome, obesity, a lipogenic phenotype, and insulin resistance, with downstream effects that may promote cancer. This mechanism is more evident if previous energy requirements are fulfilled. AMPK, adenosine monophosphate kinase; AMP, adenosine monophosphate; ATP, adenosine triphosphate.

What we propose here, based on the published literature, is that fructose can be metabolized in two different cellular contexts:

  • Poor intracellular energy (low ATP, high AMP)

  • High intracellular energy (high ATP, low AMP).

In the first situation, fructose would mainly follow the fructolytic pathway, merging the glycolytic pathway at the level of glyceraldehydes-3-phosphate and producing energy. In the second situation, with full ATP coffers, fructose would mainly be used in the lipogenic pathway, as proposed in Figure 5.

Figure 5 hypothesizes, for the first time, a contextual-dependent metabolism for fructose. A consequence of this proposed mechanism of action would be that administration of fructose under hypocaloric conditions would not induce the lipogenic pathway. Stansbie et al.162 showed that under fasting conditions, an overload of fructose generated double the amount of lactic acid compared to a non-fasted state. This suggests that fructose is mainly metabolized under hypocaloric conditions in the fructolytic pathway.

Koo et al.163 have shown that high fructose intake induced the over-expression of certain genes, such as fructokinase and aldolase B. This is logical regarding the need for these enzymes for fructolysis. However, other enzymes related to glycolysis were also over-expressed, like phosphofructokinase. Finally, there was a substantial over-expression of ChREBP. It is interesting to note that ChREBP mRNA and protein are significantly elevated in colon cancer cells compared to the normal colon, and their expression is positively associated with advanced stages of cancer.45

Other effects of fructose

Two other effects of fructose also can stimulate carcinogenesis. The first is its production of reactive oxygen species. Fructose utilization generates 100-fold more reactive oxygen species than glucose, thus, creating oxidative stress that can lead to necroinflamation,164 and oxidative stress can lead to carcinogenesis.165 The second effect is a temporary reduction in ATP levels. An overload of dietary fructose can produce a temporary ATP reduction due to the swift action of fructokinase-1 and the slow activity of aldolase B (Fig. 1).166–168 This can create an energy shortage that may temporarily restrain the lipogenic pathway. Lipogenesis is also another path that can lead to ATP depletion.169

Fructose in diagnostics

Recently, a PET scanner has been developed to image GLUT5 in breast cancer using 6-deoxy-6-[18F] fluoro-D-fructose as a radiotracer.170 This simplifies the diagnosis of fructose-dependent tumors. Another tracer that detects GLUT5 is [99mTc] glucarate, which was proposed for detecting fructose-consuming tumors in breast cancer.171

Conclusions

Glucose is the main energy source of cancer cells. Different cancer cell lines respond with considerable variations to glucose depletion.127 In certain tumors fructose becomes a very important energy source and can even replace glucose as the main nutrient. Fructose consumption is associated with a more malignant phenotype with increased proliferation, invasion, and metastasis. Some epidemiological studies with large populations could not fully confirm that high dietary fructose intake increases cancer risk.129 However, this does not mean fructose has no role in cancer development. At the clinical level, fructose indirectly participates in cancer through obesity and metabolic syndrome. At the molecular level, fructose is essential in developing an aggressive cancer phenotype in some tumors.

Clinical implications

There are several instances whereby targeting GLUT5 may be a valuable approach for clinical treatments. One of these may be in the treatment of gliomas. Gliomas have a very poor prognosis, and no effective treatment has yet been developed. Therefore, finding a new way to address this disease would be a precious tool for fighting this disease. Evidence shows that, at least in the laboratory, targeting fructose metabolism could represent an interesting add-on to conventional treatments of gliomas.60,82 Clear cell renal carcinoma over-expresses GLUT5, which represents another case in which GLUT5 targeting may improve results.64

One of the objections to the importance of fructose as an alternative energy source for cancer is the low levels of fructose in blood. This objection is invalid in colorectal cancer because fructose concentration is much higher in intestinal circulation. Moreover, as noted above, fructose concentrations can increase several folds in the bloodstream after ingestion and are increased in some types of cancer.91,92,95 Another point is that fructose can be produced endogenously from glucose through the polyol pathway, as found in diabetic patients. This pathway can convert up to 50% of glucose into fructose.172–174 The polyol pathway plays a role in cancer. The gene aldo-keto-reductase-1-member-B1, which codes for one of the two enzymes that participate in the polyol pathway, correlates with epithelial-mesenchymal transition in lung cancer patients and a colon cancer mouse model.175,176

Dietary effects

The effects of fructose in the diet deserve some more consideration. As noted above, the concept of restriction of dietary fructose to reduce cancer risk is controversial. Some researchers have suggested excessive intake of dietary fructose is an increased risk factor for cancer or related syndromes that promote cancer,17,75–77,102,128,142–148,177,178 while others could not confirm increased cancer risk in other population studies, even examining the incidence of disease in the same types of tissue.77,129–131,149,150,179 Despite this controversy, it is clear that fructose is associated with factors that indirectly have strong influences on carcinogenesis, such as obesity154,180 metabolic stress,174 diabetes155,181 and pro-inflammatory effects.182 Section “C” above proves that fructose and its metabolism and transport can promote cancer cell growth and metastasis. Therefore, decreasing dietary fructose intake should be considered as part of a cancer prevention scheme. It is unclear why population-based studies have such different and controversial results. This could be due to study group size, other parameters not controlled for, and sometimes due to differences in the types of cancer surveyed. As noted above (Fig. 5), context-dependent use of fructose may occur, and the use of fructose for metabolic activities that promote cancer may only occur in cases where high energy content is already present in cells. This could undoubtedly confuse the statistical analysis of human populations. Certain subfractions of the population may “respond” to fructose consumption more than others for this and possibly other reasons. A possibility is that GLUT5 expression levels may respond more to fructose in some parts of the population than others, possibly due to context-dependent effects. It may be that determining GLUT5 expression in individual patients could indicate a greater role of fructose in those individuals, allowing more personalized medical treatment, including fructose restriction. However, despite all the evidence on the importance of fructose metabolism in cancer, determining GLUT5 expression in cancer cells has not yet entered standard oncology practice. Targeting fructose metabolism is not part of mainstream treatments.

Related treatments, such as energy restriction, have been examined experimentally. Although successful in the laboratory, the energy restriction treatment as a stand-alone therapy with 2-deoxyglucose could not be introduced at the bedside.183 The reasons are high dose requirement, toxicity,184 and poor patient compliance. We also think that one possible explanation for this failure is the replacement of glucose with fructose as an energy source. As noted above, 2-deoxyglucose blocks glucose metabolism; however, it failed in tests as an anti-cancer drug.79 Fructose may be an alternative energy source when glucose levels are low.80 While elevated fructose metabolism by some cancer cells is not only an energetic matter, it is worth noting that 2-deoxyglucose can also decrease fructolysis.185,186 However, it is unclear whether this effect can be exploited in humans and if 2-deoxyglucose can inhibit the utilization of both fructose and glucose in tumors.

We suggest it is a mistake to consider fructose as a lonely player in cancer. The entire team of fructose, GLUT5, and KHK must be viewed as a group of pro-cancer drivers. However, GLUT5 and KHK overexpression are the product of increased fructose presence. Furthermore, each of them can independently participate in tumor progression.

Benefits of fructose restriction

We think there are beneficial effects of restriction of fructose in the diet. Western diets and sweetened beverages contain very high levels of fructose. Additionally, high fructose consumption parallels obesity incidence.187 High fructose in the diet also correlates with the progression of hepatocellular carcinoma.188 In mice predisposed to develop intestinal tumors, modest levels of high fructose corn syrup substantially increase tumor size and grade even in the absence of obesity.121 Overall, these results suggest that fructose could have an important effect on cancer patient’s diets. However, further studies are needed to demonstrate the beneficial effect of fructose restriction in humans in specific cancer types.187

Fructose consumption has experienced a significant surge since the 1960s, primarily attributed to its prevalence in high-fructose corn syrup found in soft drinks and processed foods. The relationship between high fructose intake and certain diseases remains incompletely understood. However, it is established that:

  • Increased fructose consumption can lead to weight gain and obesity through elevated lipogenesis and a surplus of calories;

  • This predisposes individuals to metabolic syndrome;

  • It also contributes to insulin resistance and diabetes;

  • Additionally, it can induce hepatic inflammation accompanied by heightened production of reactive oxygen species;

  • Ultimately, these factors collectively increase the risk of cancer.

From a molecular perspective, fructose serves as an energy source and a signaling molecule with pro-tumoral properties.

In light of this, we propose that the effects of fructose are contingent upon cellular energy availability, as illustrated in Figure 5 and Figure 6. In this hypothesis, fructolysis under conditions of low energy availability would have minimal or no impact on tumors. Conversely, fructose metabolism favors lipogenesis under sufficient energy availability, thereby promoting tumor growth.

Context-dependent metabolism of fructose.
Fig. 6  Context-dependent metabolism of fructose.

This hypothesis could elucidate the discrepancies observed in population studies and the heightened cancer risk associated with obesity and metabolic syndrome.

Furthermore, certain tumors become highly dependent on fructose, displaying a more malignant phenotype. Consequently, this finding has practical implications: patients with overexpression of the glucose transporter GLUT5 may benefit from interventions targeting fructose metabolism. Therefore, further research, including well-designed clinical trials utilizing personalized medicine, should be pursued to ascertain whether GLUT5 is overexpressed in patients’ tumors.

Abbreviations

ACC: 

Acetyl CoA Carboxylase

ADP: 

adenosine diphosphate

AMPK: 

adenosine monophosphate kinase

ATP: 

adenosine triphosphate

ChREBP: 

Carbohydrate Responsive Element Binding Protein

HIF-1α: 

hypoxia-inducible factor 1-alpha

IMP: 

inositol monophosphate

KHK: 

Ketohexokinase

PET: 

positron emission tomography

PFK: 

phosphofructokinase

RR: 

relative risk

SREBP1: 

Sterol regulatory element binding protein-1

Declarations

Acknowledgement

None to declare.

Funding

This research received no external funding.

Conflict of interest

The authors declare no conflict of interests.

Authors’ contributions

Review conception (TK); investigation, resources, writing-review, and editing (TK and LF).

References

  1. Sandulache VC, Ow TJ, Pickering CR, Frederick MJ, Zhou G, Fokt I, et al. Glucose, not glutamine, is the dominant energy source required for proliferation and survival of head and neck squamous carcinoma cells. Cancer 2011;117(13):2926-2938 View Article PubMed/NCBI
  2. Han J, Zhang L, Guo H, Wysham WZ, Roque DR, Willson AK, et al. Glucose promotes cell proliferation, glucose uptake and invasion in endometrial cancer cells via AMPK/mTOR/S6 and MAPK signaling. Gynecol Oncol 2015;138(3):668-675 View Article PubMed/NCBI
  3. Yuneva M. Finding an “Achilles’ heel” of cancer: the role of glucose and glutamine metabolism in the survival of transformed cells. Cell Cycle 2008;7(14):2083-2089 View Article PubMed/NCBI
  4. Ahmed OAK, Sibuyi NRS, Fadaka AO, Maboza E, Olivier A, Madiehe AM, et al. Prospects of Using Gum Arabic Silver Nanoparticles in Toothpaste to Prevent Dental Caries. Pharmaceutics 2023;15(3):871 View Article PubMed/NCBI
  5. Salamon S, Podbregar E, Kubatka P, Büsselberg D, Caprnda M, Opatrilova R, et al. Glucose Metabolism in Cancer and Ischemia: Possible Therapeutic Consequences of the Warburg Effect. Nutr Cancer 2017;69(2):177-183 View Article PubMed/NCBI
  6. Adekola K, Rosen ST, Shanmugam M. Glucose transporters in cancer metabolism. Curr Opin Oncol 2012;24(6):650-654 View Article PubMed/NCBI
  7. Yamamoto T, Seino Y, Fukumoto H, Koh G, Yano H, Inagaki N, et al. Over-expression of facilitative glucose transporter genes in human cancer. Biochem Biophys Res Commun 1990;170(1):223-230 View Article PubMed/NCBI
  8. Brown RS, Leung JY, Kison PV, Zasadny KR, Flint A, Wahl RL. Glucose transporters and FDG uptake in untreated primary human non-small cell lung cancer. J Nucl Med 1999;40(4):556-565 PubMed/NCBI
  9. Duhaylongsod FG, Lowe VJ, Patz EF, Vaughn AL, Coleman RE, Wolfe WG. Lung tumor growth correlates with glucose metabolism measured by fluoride-18 fluorodeoxyglucose positron emission tomography. Ann Thorac Surg 1995;60(5):1348-1352 View Article PubMed/NCBI
  10. Bailey DL, Townsend DW, Vail PE, Maisey MN. Positron Emission Tomography. London: Springer; 2003
  11. Hirayama A, Kami K, Sugimoto M, Sugawara M, Toki N, Onozuka H, et al. Quantitative metabolome profiling of colon and stomach cancer microenvironment by capillary electrophoresis time-of-flight mass spectrometry. Cancer Res 2009;69(11):4918-4925 View Article PubMed/NCBI
  12. Urasaki Y, Heath L, Xu CW. Coupling of glucose deprivation with impaired histone H2B monoubiquitination in tumors. PLoS One 2012;7(5):e36775 View Article PubMed/NCBI
  13. Laussel C, Léon S. Cellular toxicity of the metabolic inhibitor 2-deoxyglucose and associated resistance mechanisms. Biochem Pharmacol 2020;182:114213 View Article PubMed/NCBI
  14. Pajak B, Siwiak E, Sołtyka M, Priebe A, Zieliński R, Fokt I, et al. 2-Deoxy-d-Glucose and Its Analogs: From Diagnostic to Therapeutic Agents. Int J Mol Sci 2019;21(1):234 View Article PubMed/NCBI
  15. Nakagawa T, Lanaspa MA, Millan IS, Fini M, Rivard CJ, Sanchez-Lozada LG, et al. Fructose contributes to the Warburg effect for cancer growth. Cancer Metab 2020;8:16 View Article PubMed/NCBI
  16. Monzavi-Karbassi B, Hine RJ, Stanley JS, Ramani VP, Carcel-Trullols J, Whitehead TL, et al. Fructose as a carbon source induces an aggressive phenotype in MDA-MB-468 breast tumor cells. Int J Oncol 2010;37(3):615-622 View Article PubMed/NCBI
  17. Liu H, Heaney AP. Refined fructose and cancer. Expert Opin Ther Targets 2011;15(9):1049-1059 View Article PubMed/NCBI
  18. Fan X, Liu H, Liu M, Wang Y, Qiu L, Cui Y. Increased utilization of fructose has a positive effect on the development of breast cancer. PeerJ 2017;5:e3804 View Article PubMed/NCBI
  19. Sweeney MJ, Ashmore J, Morris HP, Weber G. Comparative biochemistry hepatomas. iv. isotope studies of glucose and fructose metabolism in liver tumors of different growth rates. Cancer Res 1963;23:995-1002 PubMed/NCBI
  20. Koltai T, Reshkin S, Baltasar F, Fliegel L. Prostate Cancer Metabolism. Amsterdam: Elsevier; 2021, 394
  21. Jadvar H. Is There Use for FDG-PET in Prostate Cancer?. Semin Nucl Med 2016;46(6):502-506 View Article PubMed/NCBI
  22. Takahashi N, Inoue T, Lee J, Yamaguchi T, Shizukuishi K. The roles of PET and PET/CT in the diagnosis and management of prostate cancer. Oncology 2007;72(3-4):226-233 View Article PubMed/NCBI
  23. Li R, Ravizzini GC, Gorin MA, Maurer T, Eiber M, Cooperberg MR, et al. The use of PET/CT in prostate cancer. Prostate Cancer Prostatic Dis 2018;21(1):4-21 View Article PubMed/NCBI
  24. Douard V, Choi HI, Elshenawy S, Lagunoff D, Ferraris RP. Developmental reprogramming of rat GLUT5 requires glucocorticoid receptor translocation to the nucleus. J Physiol 2008;586(15):3657-3673 View Article PubMed/NCBI
  25. Salas-Burgos A, Iserovich P, Zuniga F, Vera JC, Fischbarg J. Predicting the three-dimensional structure of the human facilitative glucose transporter glut1 by a novel evolutionary homology strategy: insights on the molecular mechanism of substrate migration, and binding sites for glucose and inhibitory molecules. Biophys J 2004;87(5):2990-2999 View Article PubMed/NCBI
  26. Reinicke K, Sotomayor P, Cisterna P, Delgado C, Nualart F, Godoy A. Cellular distribution of Glut-1 and Glut-5 in benign and malignant human prostate tissue. J Cell Biochem 2012;113(2):553-562 View Article PubMed/NCBI
  27. Carreño DV, Corro NB, Cerda-Infante JF, Echeverría CE, Asencio-Barría CA, Torres-Estay VA, et al. Dietary fructose promotes prostate cancer growth. Cancer Res 2021;81(11):2824-2832 View Article PubMed/NCBI
  28. Burant CF, Takeda J, Brot-Laroche E, Bell GI, Davidson NO. Fructose transporter in human spermatozoa and small intestine is GLUT5. J Biol Chem 1992;267(21):14523-14526 PubMed/NCBI
  29. Thorens B, Mueckler M. Glucose transporters in the 21st Century. Am J Physiol Endocrinol Metab 2010;298(2):E141-E145 View Article PubMed/NCBI
  30. McGrane MM. Carbohydrate Metabolism. Missouri: Elsevier; 2006
  31. Tappy L, Lê KA. Metabolic effects of fructose and the worldwide increase in obesity. Physiol Rev 2010;90(1):23-46 View Article PubMed/NCBI
  32. Mayes PA. Intermediary metabolism of fructose. Am J Clin Nutr 1993;58(5 Suppl):754S-765S View Article PubMed/NCBI
  33. Schaefer EJ, Gleason JA, Dansinger ML. Dietary fructose and glucose differentially affect lipid and glucose homeostasis. J Nutr 2009;139(6):1257S-1262S View Article PubMed/NCBI
  34. Vos MB, McClain CJ. Fructose takes a toll. Hepatology 2009;50(4):1004-1006 View Article PubMed/NCBI
  35. Campos VC, Tappy L. Physiological handling of dietary fructose-containing sugars: implications for health. Int J Obes (Lond) 2016;40(Suppl 1):S6-11 View Article PubMed/NCBI
  36. Watford M. Small amounts of dietary fructose dramatically increase hepatic glucose uptake through a novel mechanism of glucokinase activation. Nutr Rev 2002;60(8):253-257 View Article PubMed/NCBI
  37. Eggleston LV, Woods HF. Activation of liver pyruvate kinase by fructose-1-phosphate. FEBS Lett 1970;6(1):43-45 View Article PubMed/NCBI
  38. Ferré P, Foufelle F. Hepatic steatosis: a role for de novo lipogenesis and the transcription factor SREBP-1c. Diabetes Obes Metab 2010;12(Suppl 2):83-92 View Article PubMed/NCBI
  39. Huang WC, Li X, Liu J, Lin J, Chung LW. Activation of androgen receptor, lipogenesis, and oxidative stress converged by SREBP-1 is responsible for regulating growth and progression of prostate cancer cells. Mol Cancer Res 2012;10(1):133-142 View Article PubMed/NCBI
  40. Bao J, Zhu L, Zhu Q, Su J, Liu M, Huang W. SREBP-1 is an independent prognostic marker and promotes invasion and migration in breast cancer. Oncol Lett 2016;12(4):2409-2416 View Article PubMed/NCBI
  41. Yang Yu, Morin PJ, Han WF, Chen T, Bornman DM, Gabrielson EW, et al. Regulation of fatty acid synthase expression in breast cancer by sterol regulatory element binding protein-1c. Exp Cell Res 2003;282(2):132-137 View Article PubMed/NCBI
  42. Li C, Yang W, Zhang J, Zheng X, Yao Y, Tu K, et al. SREBP-1 has a prognostic role and contributes to invasion and metastasis in human hepatocellular carcinoma. Int J Mol Sci 2014;15(5):7124-7138 View Article PubMed/NCBI
  43. Guo D, Reinitz F, Youssef M, Hong C, Nathanson D, Akhavan D, et al. An LXR agonist promotes glioblastoma cell death through inhibition of an EGFR/AKT/SREBP-1/LDLR-dependent pathway. Cancer Discov 2011;1(5):442-456 View Article PubMed/NCBI
  44. Guo D, Bell EH, Mischel P, Chakravarti A. Targeting SREBP-1-driven lipid metabolism to treat cancer. Curr Pharm Des 2014;20(15):2619-2626 View Article PubMed/NCBI
  45. Lei Y, Zhou S, Hu Q, Chen X, Gu J. Carbohydrate response element binding protein (ChREBP) correlates with colon cancer progression and contributes to cell proliferation. Sci Rep 2020;10(1):4233 View Article PubMed/NCBI
  46. Tong X, Zhao F, Mancuso A, Gruber JJ, Thompson CB. The glucose-responsive transcription factor ChREBP contributes to glucose-dependent anabolic synthesis and cell proliferation. Proc Natl Acad Sci U S A 2009;106(51):21660-21665 View Article PubMed/NCBI
  47. Iizuka K. The transcription factor carbohydrate-response element-binding protein (ChREBP): A possible link between metabolic disease and cancer. Biochim Biophys Acta Mol Basis Dis 2017;1863(2):474-485 View Article PubMed/NCBI
  48. Airley RE, McHugh P, Evans AR, Harris B, Winchester L, Buffa FM, et al. Role of carbohydrate response element-binding protein (ChREBP) in generating an aerobic metabolic phenotype and in breast cancer progression. Br J Cancer 2014;110(3):715-723 View Article PubMed/NCBI
  49. Wang XL, Wen XF, Li RB, Liu B, Qiu GM, Wen JL, et al. Chrebp regulates the transcriptional activity of androgen receptor in prostate cancer. Tumour Biol 2014;35(8):8143-8148 View Article PubMed/NCBI
  50. Barone S, Fussell SL, Singh AK, Lucas F, Xu J, Kim C, et al. Slc2a5 (Glut5) is essential for the absorption of fructose in the intestine and generation of fructose-induced hypertension. J Biol Chem 2009;284(8):5056-5066 View Article PubMed/NCBI
  51. Chen WL, Wang YY, Zhao A, Xia L, Xie G, Su M, et al. Enhanced Fructose Utilization Mediated by SLC2A5 Is a Unique Metabolic Feature of Acute Myeloid Leukemia with Therapeutic Potential. Cancer Cell 2016;30(5):779-791 View Article PubMed/NCBI
  52. Jin C, Gong X, Shang Y. GLUT5 increases fructose utilization in ovarian cancer. Onco Targets Ther 2019;12:5425-5436 View Article PubMed/NCBI
  53. Medina Villaamil V, Aparicio Gallego G, Valbuena Rubira L, García Campelo R, Valladares-Ayerbes M, Grande Pulido E, et al. Fructose transporter GLUT5 expression in clear renal cell carcinoma. Oncol Rep 2011;25(2):315-323 View Article PubMed/NCBI
  54. Zamora-León SP, Golde DW, Concha II, Rivas CI, Delgado-López F, Baselga J, et al. Expression of the fructose transporter GLUT5 in human breast cancer. Proc Natl Acad Sci U S A 1996;93(5):1847-1852 View Article PubMed/NCBI
  55. Sasaki A, Yamaguchi H, Horikoshi Y, Tanaka G, Nakazato Y. Expression of glucose transporter 5 by microglia in human gliomas. Neuropathol Appl Neurobiol 2004;30(5):447-455 View Article PubMed/NCBI
  56. Weng Y, Fan X, Bai Y, Wang S, Huang H, Yang H, et al. SLC2A5 promotes lung adenocarcinoma cell growth and metastasis by enhancing fructose utilization. Cell Death Discov 2018;4:38 View Article PubMed/NCBI
  57. Włodarczyk J, Włodarczyk M, Zielińska M, Jędrzejczak B, Dziki Ł, Fichna J. Blockade of fructose transporter protein GLUT5 inhibits proliferation of colon cancer cells: proof of concept for a new class of anti-tumor therapeutics. Pharmacol Rep 2021;73(3):939-945 View Article PubMed/NCBI
  58. Yan TY, Naren DL, Gong YP. The Roles of Glut5 in Imatinib Resistance in the Ph+ Acute Lymphoblastic Leukemia Cell. Sichuan Da Xue Xue Bao Yi Xue Ban 2017;48(3):389-393 PubMed/NCBI
  59. Fang JH, Chen JY, Zheng JL, Zeng HX, Chen JG, Wu CH, et al. Fructose metabolism in tumor endothelial cells promotes angiogenesis by activating AMPK signaling and mitochondrial respiration. Cancer Res 2023;83(8):1249-1263 View Article PubMed/NCBI
  60. Su C, Li H, Gao W. GLUT5 increases fructose utilization and promotes tumor progression in glioma. Biochem Biophys Res Commun 2018;500(2):462-469 View Article PubMed/NCBI
  61. Shen Z, Li Z, Liu Y, Li Y, Feng X, Zhan Y, et al. GLUT5-KHK axis-mediated fructose metabolism drives proliferation and chemotherapy resistance of colorectal cancer. Cancer Lett 2022;534:215617 View Article PubMed/NCBI
  62. Medina Villaamil V, Aparicio Gallego G, Santamarina Caínzos I, Valladares-Ayerbes M, Antón Aparicio LM. Searching for Hif1-α interacting proteins in renal cell carcinoma. Clin Transl Oncol 2012;14(9):698-708 View Article PubMed/NCBI
  63. Huang X, Fang J, Lai W, Hu Y, Li L, Zhong Y, et al. IL-6/STAT3 Axis Activates Glut5 to Regulate Fructose Metabolism and Tumorigenesis. Int J Biol Sci 2022;18(9):3668-3675 View Article PubMed/NCBI
  64. Jin X, Liang Y, Liu D, Luo Q, Cai L, Wu J, et al. An essential role for GLUT5-mediated fructose utilization in exacerbating the malignancy of clear cell renal cell carcinoma. Cell Biol Toxicol 2019;35(5):471-483 View Article PubMed/NCBI
  65. Chan KK, Chan JY, Chung KK, Fung KP. Inhibition of cell proliferation in human breast tumor cells by antisense oligonucleotides against facilitative glucose transporter 5. J Cell Biochem 2004;93(6):1134-1142 View Article PubMed/NCBI
  66. Groenendyk J, Stoletov K, Paskevicius T, Li W, Dai N, Pujol M, et al. Loss of the fructose transporter SLC2A5 inhibits cancer cell migration. Front Cell Dev Biol 2022;10:896297 View Article PubMed/NCBI
  67. Enzmann H, Ohlhauser D, Dettler T, Bannasch P. Enhancement of hepatocarcinogenesis in rats by dietary fructose. Carcinogenesis 1989;10(7):1247-1252 View Article PubMed/NCBI
  68. Enzmann H, Dettler T, Ohlhauser D, Stumpf H, Bannasch P. Dietary fructose enhances the development of atypical acinar cell nodules in the pancreas of rats pretreated with N-nitrosomorpholine. Arch Geschwulstforsch 1990;60(4):283-287 PubMed/NCBI
  69. Liang RJ, Taylor S, Nahiyaan N, Song J, Murphy CJ, Dantas E, et al. GLUT5 (SLC2A5) enables fructose-mediated proliferation independent of ketohexokinase. Cancer Metab 2021;9(1):12 View Article PubMed/NCBI
  70. Jiang Y, Pan Y, Rhea PR, Tan L, Gagea M, Cohen L, et al. A Sucrose-Enriched Diet Promotes Tumorigenesis in Mammary Gland in Part through the 12-Lipoxygenase Pathway. Cancer Res 2016;76(1):24-29 View Article PubMed/NCBI
  71. Ozawa T, Maehara N, Kai T, Arai S, Miyazaki T. Dietary fructose-induced hepatocellular carcinoma development manifested in mice lacking apoptosis inhibitor of macrophage (AIM). Genes Cells 2016;21(12):1320-1332 View Article PubMed/NCBI
  72. Kuehm LM, Khojandi N, Piening A, Klevorn LE, Geraud SC, McLaughlin NR, et al. Fructose Promotes Cytoprotection in Melanoma Tumors and Resistance to Immunotherapy. Cancer Immunol Res 2021;9(2):227-238 View Article PubMed/NCBI
  73. Taylor SR, Ramsamooj S, Liang RJ, Katti A, Pozovskiy R, Vasan N, et al. Dietary fructose improves intestinal cell survival and nutrient absorption. Nature 2021;597(7875):263-267 View Article PubMed/NCBI
  74. Michaud DS, Liu S, Giovannucci E, Willett WC, Colditz GA, Fuchs CS. Dietary sugar, glycemic load, and pancreatic cancer risk in a prospective study. J Natl Cancer Inst 2002;94(17):1293-1300 View Article PubMed/NCBI
  75. Larsson SC, Bergkvist L, Wolk A. Consumption of sugar and sugar-sweetened foods and the risk of pancreatic cancer in a prospective study. Am J Clin Nutr 2006;84(5):1171-1176 View Article PubMed/NCBI
  76. Nöthlings U, Murphy SP, Wilkens LR, Henderson BE, Kolonel LN. Dietary glycemic load, added sugars, and carbohydrates as risk factors for pancreatic cancer: the Multiethnic Cohort Study. Am J Clin Nutr 2007;86(5):1495-1501 View Article PubMed/NCBI
  77. Aune D, Chan DSM, Vieira AR, Navarro Rosenblatt DA, Vieira R, Greenwood DC, et al. Dietary fructose, carbohydrates, glycemic indices and pancreatic cancer risk: a systematic review and meta-analysis of cohort studies. Ann Oncol 2012;23(10):2536-2546 View Article PubMed/NCBI
  78. Koltai T. Cancer: fundamentals behind pH targeting and the double-edged approach. Onco Targets Ther 2016;9:6343-6360 View Article PubMed/NCBI
  79. Tannock IF, Guttman P, Rauth AM. Failure of 2-deoxy-D-glucose and 5-thio-D-glucose to kill hypoxic cells of two murine tumors. Cancer Res 1983;43(3):980-983 PubMed/NCBI
  80. Weng Y, Zhu J, Chen Z, Fu J, Zhang F. Fructose fuels lung adenocarcinoma through GLUT5. Cell Death Dis 2018;9(5):557 View Article PubMed/NCBI
  81. Godoy A, Ulloa V, Rodríguez F, Reinicke K, Yañez AJ, García Mde L, et al. Differential subcellular distribution of glucose transporters GLUT1-6 and GLUT9 in human cancer: ultrastructural localization of GLUT1 and GLUT5 in breast tumor tissues. J Cell Physiol 2006;207(3):614-627 View Article PubMed/NCBI
  82. Gao W, Li N, Li Z, Xu J, Su C. Ketohexokinase is involved in fructose utilization and promotes tumor progression in glioma. Biochem Biophys Res Commun 2018;503(3):1298-1306 View Article PubMed/NCBI
  83. Bu P, Chen KY, Xiang K, Johnson C, Crown SB, Rakhilin N, et al. Aldolase B-Mediated Fructose Metabolism Drives Metabolic Reprogramming of Colon Cancer Liver Metastasis. Cell Metab 2018;27(6):1249-1262.e4 View Article PubMed/NCBI
  84. Liu H, Huang D, McArthur DL, Boros LG, Nissen N, Heaney AP. Fructose induces transketolase flux to promote pancreatic cancer growth. Cancer Res 2010;70(15):6368-6376 View Article PubMed/NCBI
  85. Shen CN, Hsieh CC, Li WS, Hsiao M. Therapeutic implication of identifying pancreatic cancer stem cells possessing fructose metabolic signature. Cancer Res 2016;76(14 Suppl):2489 View Article
  86. Carreño D, Corro N, Arredondo M, Navarro C, Torres V, Montecinos V, et al. Role of fructose in prostate cancer. Cancer Res 2017;77(13 Suppl):448 View Article
  87. Kim J, Kang J, Kang YL, Woo J, Kim Y, Huh J, et al. Ketohexokinase-A acts as a nuclear protein kinase that mediates fructose-induced metastasis in breast cancer. Nat Commun 2020;11(1):5436 View Article PubMed/NCBI
  88. Chávez-Rodríguez L, Escobedo-Calvario A, Simoni-Nieves A, Bucio L, Souza V, Labra RM, et al. Fructose diet induces a metabolic reprogramming to enhance tumor aggressiveness. Ann Hepatol 2022;27:100628 View Article
  89. Shu R, David ES, Ferraris RP. Dietary fructose enhances intestinal fructose transport and GLUT5 expression in weaning rats. Am J Physiol 1997;272(3 Pt 1):G446-G453 View Article PubMed/NCBI
  90. Jiang L, David ES, Espina N, Ferraris RP. GLUT-5 expression in neonatal rats: crypt-villus location and age-dependent regulation. Am J Physiol Gastrointest Liver Physiol 2001;281(3):G666-G674 View Article PubMed/NCBI
  91. Patel C, Sugimoto K, Douard V, Shah A, Inui H, Yamanouchi T, et al. Effect of dietary fructose on portal and systemic serum fructose levels in rats and in KHK-/- and GLUT5-/- mice. Am J Physiol Gastrointest Liver Physiol 2015;309(9):G779-G790 View Article PubMed/NCBI
  92. Wahjudi PN, Patterson ME, Lim S, Yee JK, Mao CS, Lee WN. Measurement of glucose and fructose in clinical samples using gas chromatography/mass spectrometry. Clin Biochem 2010;43(1-2):198-207 View Article PubMed/NCBI
  93. Macdonald I, Keyser A, Pacy D. Some effects, in man, of varying the load of glucose, sucrose, fructose, or sorbitol on various metabolites in blood. Am J Clin Nutr 1978;31(8):1305-1311 View Article PubMed/NCBI
  94. Le MT, Frye RF, Rivard CJ, Cheng J, McFann KK, Segal MS, et al. Effects of high-fructose corn syrup and sucrose on the pharmacokinetics of fructose and acute metabolic and hemodynamic responses in healthy subjects. Metabolism 2012;61(5):641-651 View Article PubMed/NCBI
  95. Hui H, Huang D, McArthur D, Nissen N, Boros LG, Heaney AP. Direct spectrophotometric determination of serum fructose in pancreatic cancer patients. Pancreas 2009;38(6):706-712 View Article PubMed/NCBI
  96. Mueckler M, Thorens B. The SLC2 (GLUT) family of membrane transporters. Mol Aspects Med 2013;34(2-3):121-138 View Article PubMed/NCBI
  97. Nomura N, Verdon G, Kang HJ, Shimamura T, Nomura Y, Sonoda Y, et al. Structure and mechanism of the mammalian fructose transporter GLUT5. Nature 2015;526(7573):397-401 View Article PubMed/NCBI
  98. Shepherd PR, Gibbs EM, Wesslau C, Gould GW, Kahn BB. Human small intestine facilitative fructose/glucose transporter (GLUT5) is also present in insulin-responsive tissues and brain. Investigation of biochemical characteristics and translocation. Diabetes 1992;41(10):1360-1365 View Article PubMed/NCBI
  99. Shu R, David ES, Ferraris RP. Luminal fructose modulates fructose transport and GLUT-5 expression in small intestine of weaning rats. Am J Physiol 1998;274(2):G232-G239 View Article PubMed/NCBI
  100. Gouyon F, Onesto C, Dalet V, Pages G, Leturque A, Brot-Laroche E. Fructose modulates GLUT5 mRNA stability in differentiated Caco-2 cells: role of cAMP-signalling pathway and PABP (polyadenylated-binding protein)-interacting protein (Paip) 2. Biochem J 2003;375(Pt 1):167-174 View Article PubMed/NCBI
  101. Hajduch E, Litherland GJ, Turban S, Brot-Laroche E, Hundal HS. Insulin regulates the expression of the GLUT5 transporter in L6 skeletal muscle cells. FEBS Lett 2003;549(1-3):77-82 View Article PubMed/NCBI
  102. Hannou SA, Haslam DE, McKeown NM, Herman MA. Fructose metabolism and metabolic disease. J Clin Invest 2018;128(2):545-555 View Article PubMed/NCBI
  103. Matosin-Matekalo M, Mesonero JE, Laroche TJ, Lacasa M, Brot-Laroche E. Glucose and thyroid hormone co-regulate the expression of the intestinal fructose transporter GLUT5. Biochem J 1999;339(2):233-239 PubMed/NCBI
  104. Wood IS, Wang B, Lorente-Cebrián S, Trayhurn P. Hypoxia increases expression of selective facilitative glucose transporters (GLUT) and 2-deoxy-D-glucose uptake in human adipocytes. Biochem Biophys Res Commun 2007;361(2):468-473 View Article PubMed/NCBI
  105. Hamann I, Krys D, Glubrecht D, Bouvet V, Marshall A, Vos L, et al. Expression and function of hexose transporters GLUT1, GLUT2, and GLUT5 in breast cancer-effects of hypoxia. FASEB J 2018;32(9):5104-5118 View Article PubMed/NCBI
  106. Chen WL, Jin X, Wang M, Liu D, Luo Q, Tian H, et al. GLUT5-mediated fructose utilization drives lung cancer growth by stimulating fatty acid synthesis and AMPK/mTORC1 signaling. JCI Insight 2020;5(3):131596 View Article PubMed/NCBI
  107. Ramzy GM, Boschung L, Koessler T, Delucinge-Vivier C, Docquier M, McKee TA, et al. FOLFOXIRI Resistance Induction and Characterization in Human Colorectal Cancer Cells. Cancers (Basel) 2022;14(19):4812 View Article PubMed/NCBI
  108. Park GB, Jeong JY, Kim D. GLUT5 regulation by AKT1/3-miR-125b-5p downregulation induces migratory activity and drug resistance in TLR-modified colorectal cancer cells. Carcinogenesis 2020;41(10):1329-1340 View Article PubMed/NCBI
  109. Yang J, Dong C, Wu J, Liu D, Luo Q, Jin X. Fructose utilization enhanced by GLUT5 promotes lung cancer cell migration via activating glycolysis/AKT pathway. Clin Transl Oncol 2023;25(4):1080-1090 View Article PubMed/NCBI
  110. Raushel FM, Cleland WW. Bovine liver fructokinase: purification and kinetic properties. Biochemistry 1977;16(10):2169-2175 View Article PubMed/NCBI
  111. Raushel FM, Cleland WW. Determination of the rate-limiting steps and chemical mechanism of fructokinase by isotope exchange, isotope partitioning, and pH studies. Biochemistry 1977;16(10):2176-2181 View Article PubMed/NCBI
  112. Bonthron DT, Brady N, Donaldson IA, Steinmann B. Molecular basis of essential fructosuria: molecular cloning and mutational analysis of human ketohexokinase (fructokinase). Hum Mol Genet 1994;3(9):1627-1631 View Article PubMed/NCBI
  113. Mirtschink P, Krishnan J, Grimm F, Sarre A, Hörl M, Kayikci M, et al. HIF-driven SF3B1 induces KHK-C to enforce fructolysis and heart disease. Nature 2015;522(7557):444-449 View Article PubMed/NCBI
  114. Snaebjornsson MT, Schulze A. Non-canonical functions of enzymes facilitate cross-talk between cell metabolic and regulatory pathways. Exp Mol Med 2018;50(4):1-16 View Article PubMed/NCBI
  115. Yang X, Shao F, Shi S, Feng X, Wang W, Wang Y, et al. Prognostic Impact of Metabolism Reprogramming Markers Acetyl-CoA Synthetase 2 Phosphorylation and Ketohexokinase-A Expression in Non-Small-Cell Lung Carcinoma. Front Oncol 2019;9:1123 View Article PubMed/NCBI
  116. Deng Q, Wu M, Deng J. USP36 promotes tumor growth of non-small cell lung cancer via increasing KHK-A expression by regulating c-MYC-hnRNPH1/H2 axis. Hum Cell 2022;35(2):694-704 View Article PubMed/NCBI
  117. Tang G. Ketohexokinase and fructose metabolism promote pancreatic cancer growth by activating MAPK signaling pathway [Dissertation]. Switzerland: ETH Zurich; 2019 View Article
  118. Li X, Qian X, Peng LX, Jiang Y, Hawke DH, Zheng Y, et al. Corrigendum: A splicing switch from ketohexokinase-C to ketohexokinase-A drives hepatocellular carcinoma formation. Nat Cell Biol 2016;18(6):709 View Article PubMed/NCBI
  119. Yang J, Dowden J, Tatibouët A, Hatanaka Y, Holman GD. Development of high-affinity ligands and photoaffinity labels for the D-fructose transporter GLUT5. Biochem J 2002;367(Pt 2):533-539 View Article PubMed/NCBI
  120. Kanarek N, Petrova B, Sabatini DM. Dietary modifications for enhanced cancer therapy. Nature 2020;579(7800):507-517 View Article PubMed/NCBI
  121. Goncalves MD, Lu C, Tutnauer J, Hartman TE, Hwang SK, Murphy CJ, et al. High-fructose corn syrup enhances intestinal tumor growth in mice. Science 2019;363(6433):1345-1349 View Article PubMed/NCBI
  122. George Thompson AM, Ursu O, Babkin P, Iancu CV, Whang A, Oprea TI, et al. Discovery of a specific inhibitor of human GLUT5 by virtual screening and in vitro transport evaluation. Sci Rep 2016;6:24240 View Article PubMed/NCBI
  123. Girniene J, Tatibouët A, Sackus A, Yang J, Holman GD, Rollin P. Inhibition of the D-fructose transporter protein GLUT5 by fused-ring glyco-1,3-oxazolidin-2-thiones and -oxazolidin-2-ones. Carbohydr Res 2003;338(8):711-719 View Article PubMed/NCBI
  124. Villa-Rodriguez JA, Aydin E, Gauer JS, Pyner A, Williamson G, Kerimi A. Green and Chamomile Teas, but not Acarbose, Attenuate Glucose and Fructose Transport via Inhibition of GLUT2 and GLUT5. Mol Nutr Food Res 2017;61(12):1700566 View Article PubMed/NCBI
  125. Gauer JS, Tumova S, Lippiat JD, Kerimi A, Williamson G. Differential patterns of inhibition of the sugar transporters GLUT2, GLUT5 and GLUT7 by flavonoids. Biochem Pharmacol 2018;152:11-20 View Article PubMed/NCBI
  126. George Thompson AM, Iancu CV, Nguyen TT, Kim D, Choe JY. Inhibition of human GLUT1 and GLUT5 by plant carbohydrate products; insights into transport specificity. Sci Rep 2015;5:12804 View Article PubMed/NCBI
  127. Birsoy K, Possemato R, Lorbeer FK, Bayraktar EC, Thiru P, Yucel B, et al. Metabolic determinants of cancer cell sensitivity to glucose limitation and biguanides. Nature 2014;508(7494):108-112 View Article PubMed/NCBI
  128. Port AM, Ruth MR, Istfan NW. Fructose consumption and cancer: is there a connection?. Curr Opin Endocrinol Diabetes Obes 2012;19(5):367-374 View Article PubMed/NCBI
  129. Chan JM, Wang F, Holly EA. Sweets, sweetened beverages, and risk of pancreatic cancer in a large population-based case-control study. Cancer Causes Control 2009;20(6):835-846 View Article PubMed/NCBI
  130. Maino Vieytes CA, Taha HM, Burton-Obanla AA, Douglas KG, Arthur AE. Carbohydrate Nutrition and the Risk of Cancer. Curr Nutr Rep 2019;8(3):230-239 View Article PubMed/NCBI
  131. Higginbotham S, Zhang ZF, Lee IM, Cook NR, Giovannucci E, Buring JE, et al. Dietary glycemic load and risk of colorectal cancer in the Women’s Health Study. J Natl Cancer Inst 2004;96(3):229-233 View Article PubMed/NCBI
  132. Ludwig DS, Gortmaker SL. Programming obesity in childhood. Lancet 2004;364(9430):226-227 View Article PubMed/NCBI
  133. Rutledge AC, Adeli K. Fructose and the metabolic syndrome: pathophysiology and molecular mechanisms. Nutr Rev 2007;65(6 Pt 2):S13-S23 View Article PubMed/NCBI
  134. Havel PJ. Dietary fructose: implications for dysregulation of energy homeostasis and lipid/carbohydrate metabolism. Nutr Rev 2005;63(5):133-157 View Article PubMed/NCBI
  135. Das UN. Sucrose, fructose, glucose, and their link to metabolic syndrome and cancer. Nutrition 2015;31(1):249-257 View Article PubMed/NCBI
  136. Basciano H, Federico L, Adeli K. Fructose, insulin resistance, and metabolic dyslipidemia. Nutr Metab (Lond) 2005;2(1):5 View Article PubMed/NCBI
  137. Fried SK, Bunkin DA, Greenberg AS. Omental and subcutaneous adipose tissues of obese subjects release interleukin-6: depot difference and regulation by glucocorticoid. J Clin Endocrinol Metab 1998;83(3):847-850 View Article PubMed/NCBI
  138. Cai D, Yuan M, Frantz DF, Melendez PA, Hansen L, Lee J, et al. Local and systemic insulin resistance resulting from hepatic activation of IKK-beta and NF-kappaB. Nat Med 2005;11(2):183-190 View Article PubMed/NCBI
  139. Baena M, Sangüesa G, Dávalos A, Latasa MJ, Sala-Vila A, Sánchez RM, et al. Fructose, but not glucose, impairs insulin signaling in the three major insulin-sensitive tissues. Sci Rep 2016;6:26149 View Article PubMed/NCBI
  140. Esposito K, Chiodini P, Colao A, Lenzi A, Giugliano D. Metabolic syndrome and risk of cancer: a systematic review and meta-analysis. Diabetes Care 2012;35(11):2402-2411 View Article PubMed/NCBI
  141. Calle EE, Rodriguez C, Walker-Thurmond K, Thun MJ. Overweight, obesity, and mortality from cancer in a prospectively studied cohort of U.S. adults. N Engl J Med 2003;348(17):1625-1638 View Article PubMed/NCBI
  142. Bray GA, Nielsen SJ, Popkin BM. Consumption of high-fructose corn syrup in beverages may play a role in the epidemic of obesity. Am J Clin Nutr 2004;79(4):537-543 View Article PubMed/NCBI
  143. Bray GA. Soft drink consumption and obesity: it is all about fructose. Curr Opin Lipidol 2010;21(1):51-57 View Article PubMed/NCBI
  144. Bocarsly ME, Powell ES, Avena NM, Hoebel BG. High-fructose corn syrup causes characteristics of obesity in rats: increased body weight, body fat and triglyceride levels. Pharmacol Biochem Behav 2010;97(1):101-106 View Article PubMed/NCBI
  145. Perez-Pozo SE, Schold J, Nakagawa T, Sánchez-Lozada LG, Johnson RJ, Lillo JL. Excessive fructose intake induces the features of metabolic syndrome in healthy adult men: role of uric acid in the hypertensive response. Int J Obes (Lond) 2010;34(3):454-461 View Article PubMed/NCBI
  146. Stanhope KL, Schwarz JM, Keim NL, Griffen SC, Bremer AA, Graham JL, et al. Consuming fructose-sweetened, not glucose-sweetened, beverages increases visceral adiposity and lipids and decreases insulin sensitivity in overweight/obese humans. J Clin Invest 2009;119(5):1322-1334 View Article PubMed/NCBI
  147. Lustig RH, Mulligan K, Noworolski SM, Tai VW, Wen MJ, Erkin-Cakmak A, et al. Isocaloric fructose restriction and metabolic improvement in children with obesity and metabolic syndrome. Obesity (Silver Spring) 2016;24(2):453-460 View Article PubMed/NCBI
  148. Brinton EA. The time has come to flag and reduce excess fructose intake. Atherosclerosis 2016;253:262-264 View Article PubMed/NCBI
  149. van Buul VJ, Tappy L, Brouns FJ. Misconceptions about fructose-containing sugars and their role in the obesity epidemic. Nutr Res Rev 2014;27(1):119-130 View Article PubMed/NCBI
  150. Stanhope KL. Role of fructose-containing sugars in the epidemics of obesity and metabolic syndrome. Annu Rev Med 2012;63:329-343 View Article PubMed/NCBI
  151. Swinnen JV, Heemers H, van de Sande T, de Schrijver E, Brusselmans K, Heyns W, et al. Androgens, lipogenesis and prostate cancer. J Steroid Biochem Mol Biol 2004;92(4):273-279 View Article PubMed/NCBI
  152. Heemers H, Maes B, Foufelle F, Heyns W, Verhoeven G, Swinnen JV. Androgens stimulate lipogenic gene expression in prostate cancer cells by activation of the sterol regulatory element-binding protein cleavage activating protein/sterol regulatory element-binding protein pathway. Mol Endocrinol 2001;15(10):1817-1828 View Article PubMed/NCBI
  153. Carreño D, Corro N, Torres-Estay V, Véliz LP, Jaimovich R, Cisternas P, et al. Fructose and prostate cancer: toward an integrated view of cancer cell metabolism. Prostate Cancer Prostatic Dis 2019;22(1):49-58 View Article PubMed/NCBI
  154. Robubi A, Huber KR, Krugluger W. Extra fructose in the growth medium fuels lipogenesis of adipocytes. J Obes 2014;2014:647034 View Article PubMed/NCBI
  155. Nagai Y, Yonemitsu S, Erion DM, Iwasaki T, Stark R, Weismann D, et al. The role of peroxisome proliferator-activated receptor gamma coactivator-1 beta in the pathogenesis of fructose-induced insulin resistance. Cell Metab 2009;9(3):252-264 View Article PubMed/NCBI
  156. Dekker MJ, Su Q, Baker C, Rutledge AC, Adeli K. Fructose: a highly lipogenic nutrient implicated in insulin resistance, hepatic steatosis, and the metabolic syndrome. Am J Physiol Endocrinol Metab 2010;299(5):E685-E694 View Article PubMed/NCBI
  157. Lê KA, Faeh D, Stettler R, Ith M, Kreis R, Vermathen P, et al. A 4-wk high-fructose diet alters lipid metabolism without affecting insulin sensitivity or ectopic lipids in healthy humans. Am J Clin Nutr 2006;84(6):1374-1379 View Article PubMed/NCBI
  158. Forshee RA, Storey ML, Allison DB, Glinsmann WH, Hein GL, Lineback DR, et al. A critical examination of the evidence relating high fructose corn syrup and weight gain. Crit Rev Food Sci Nutr 2007;47(6):561-582 View Article PubMed/NCBI
  159. Gugliucci A. Fructose surges damage hepatic adenosyl-monophosphate-dependent kinase and lead to increased lipogenesis and hepatic insulin resistance. Med Hypotheses 2016;93:87-92 View Article PubMed/NCBI
  160. Lally JSV, Ghoshal S, DePeralta DK, Moaven O, Wei L, Masia R, et al. Inhibition of Acetyl-CoA Carboxylase by Phosphorylation or the Inhibitor ND-654 Suppresses Lipogenesis and Hepatocellular Carcinoma. Cell Metab 2019;29(1):174-182.e5 View Article PubMed/NCBI
  161. Woods A, Williams JR, Muckett PJ, Mayer FV, Liljevald M, Bohlooly-Y M, et al. Liver-Specific Activation of AMPK Prevents Steatosis on a High-Fructose Diet. Cell Rep 2017;18(13):3043-3051 View Article PubMed/NCBI
  162. Stansbie D, Sherriff RJ, Denton RM. Fructose load test—an in vivo screening test designed to assess pyruvate dehydrogenase activity and interconversion. J Inherit Metab Dis 1978;1(4):163-165 View Article PubMed/NCBI
  163. Koo HY, Wallig MA, Chung BH, Nara TY, Cho BH, Nakamura MT. Dietary fructose induces a wide range of genes with distinct shift in carbohydrate and lipid metabolism in fed and fasted rat liver. Biochim Biophys Acta 2008;1782(5):341-348 View Article PubMed/NCBI
  164. Lim JS, Mietus-Snyder M, Valente A, Schwarz JM, Lustig RH. The role of fructose in the pathogenesis of NAFLD and the metabolic syndrome. Nat Rev Gastroenterol Hepatol 2010;7(5):251-264 View Article PubMed/NCBI
  165. Takaki A, Yamamoto K. Control of oxidative stress in hepatocellular carcinoma: Helpful or harmful?. World J Hepatol 2015;7(7):968-979 View Article PubMed/NCBI
  166. Theytaz F, de Giorgi S, Hodson L, Stefanoni N, Rey V, Schneiter P, et al. Metabolic fate of fructose ingested with and without glucose in a mixed meal. Nutrients 2014;6(7):2632-2649 View Article PubMed/NCBI
  167. Jegatheesan P, De Bandt JP. Fructose and NAFLD: The Multifaceted Aspects of Fructose Metabolism. Nutrients 2017;9(3):230 View Article PubMed/NCBI
  168. Woods HF, Alberti KG. Dangers of intravenous fructose. Lancet 1972;2(7791):1354-1357 View Article PubMed/NCBI
  169. Ouyang X, Cirillo P, Sautin Y, McCall S, Bruchette JL, Diehl AM, et al. Fructose consumption as a risk factor for non-alcoholic fatty liver disease. J Hepatol 2008;48(6):993-999 View Article PubMed/NCBI
  170. Bouvet V, Jans HS, Wuest M, Soueidan OM, Mercer J, McEwan AJ, et al. Automated synthesis and dosimetry of 6-deoxy-6-[(18)F]fluoro-D-fructose (6-[(18)F]FDF): a radiotracer for imaging of GLUT5 in breast cancer. Am J Nucl Med Mol Imaging 2014;4(3):248-259 PubMed/NCBI
  171. Isnardi V, Clotagatide A, Bruel S, Perek N. Is [(99m)Tc]glucarate uptake mediated by fructose transporter GLUT-5?. Nucl Med Biol 2012;39(8):1226-1231 View Article PubMed/NCBI
  172. Andres-Hernando A, Johnson RJ, Lanaspa MA. Endogenous fructose production: what do we know and how relevant is it?. Curr Opin Clin Nutr Metab Care 2019;22(4):289-294 View Article PubMed/NCBI
  173. Lanaspa MA, Ishimoto T, Cicerchi C, Tamura Y, Roncal-Jimenez CA, Chen W, et al. Endogenous fructose production and fructokinase activation mediate renal injury in diabetic nephropathy. J Am Soc Nephrol 2014;25(11):2526-2538 View Article PubMed/NCBI
  174. Lanaspa MA, Ishimoto T, Li N, Cicerchi C, Orlicky DJ, Ruzycki P, et al. Endogenous fructose production and metabolism in the liver contributes to the development of metabolic syndrome. Nat Commun 2013;4:2434 View Article PubMed/NCBI
  175. Schwab A, Siddiqui A, Vazakidou ME, Napoli F, Böttcher M, Menchicchi B, et al. Polyol Pathway Links Glucose Metabolism to the Aggressiveness of Cancer Cells. Cancer Res 2018;78(7):1604-1618 View Article PubMed/NCBI
  176. Ceppi P, Schwab A. PO-166 Polyol pathway connects glucose metabolism with cancer differentiation and EMT. ESMO Open 2018;3:A86 View Article
  177. Tasevska N, Jiao L, Cross AJ, Kipnis V, Subar AF, Hollenbeck A, et al. Sugars in diet and risk of cancer in the NIH-AARP Diet and Health Study. Int J Cancer 2012;130(1):159-169 View Article PubMed/NCBI
  178. Charrez B, Qiao L, Hebbard L. The role of fructose in metabolism and cancer. Horm Mol Biol Clin Investig 2015;22(2):79-89 View Article PubMed/NCBI
  179. Giovannucci E, Rimm EB, Wolk A, Ascherio A, Stampfer MJ, Colditz GA, et al. Calcium and fructose intake in relation to risk of prostate cancer. Cancer Res 1998;58(3):442-447 PubMed/NCBI
  180. Stanhope KL, Havel PJ. Fructose consumption: considerations for future research on its effects on adipose distribution, lipid metabolism, and insulin sensitivity in humans. J Nutr 2009;139(6):1236S-1241S View Article PubMed/NCBI
  181. Elliott SS, Keim NL, Stern JS, Teff K, Havel PJ. Fructose, weight gain, and the insulin resistance syndrome. Am J Clin Nutr 2002;76(5):911-922 View Article PubMed/NCBI
  182. Veličković N, Djordjevic A, Vasiljević A, Bursać B, Milutinović DV, Matić G. Tissue-specific regulation of inflammation by macrophage migration inhibitory factor and glucocorticoids in fructose-fed Wistar rats. Br J Nutr 2013;110(3):456-465 View Article PubMed/NCBI
  183. Ishino K, Kudo M, Peng WX, Kure S, Kawahara K, Teduka K, et al. 2-Deoxy-d-glucose increases GFAT1 phosphorylation resulting in endoplasmic reticulum-related apoptosis via disruption of protein N-glycosylation in pancreatic cancer cells. Biochem Biophys Res Commun 2018;501(3):668-673 View Article PubMed/NCBI
  184. Vijayaraghavan R, Kumar D, Dube SN, Singh R, Pandey KS, Bag BC, et al. Acute toxicity and cardio-respiratory effects of 2-deoxy-D-glucose: a promising radio sensitiser. Biomed Environ Sci 2006;19(2):96-103 PubMed/NCBI
  185. Lodge JR, Harms PG, Graves CN. Effect of 2-deoxyglucose on bovine spermatozoan metabolism and preservation. J Dairy Sci 1968;51(7):1085-1090 View Article PubMed/NCBI
  186. Woodward GE, Hudson MT. The effect of 2-desoxy-D-glucose on glycolysis and respiration of tumor and normal tissues. Cancer Res 1954;14(8):599-605 PubMed/NCBI
  187. Morita M, Kudo K, Shima H, Tanuma N. Dietary intervention as a therapeutic for cancer. Cancer Sci 2021;112(2):498-504 View Article PubMed/NCBI
  188. Chávez-Rodríguez L, Escobedo-Calvario A, Salas-Silva S, Miranda-Labra RU, Bucio L, Souza V, et al. Fructose consumption and hepatocellular carcinoma promotion. Livers 2021;1:250-262 View Article
  • Gene Expression
  • pISSN 1052-2166
  • eISSN 1555-3884
Back to Top

Fructose, Another Sweet for Cancer: A Context Acting Nutrient Hypothesis

Tomas Koltai, Larry Fliegel
  • Reset Zoom
  • Download TIFF