Next Article in Journal
Insight into the Role of Angiopoietins in Ageing-Associated Diseases
Next Article in Special Issue
Profiling and Targeting of Energy and Redox Metabolism in Grade 2 Bladder Cancer Cells with Different Invasiveness Properties
Previous Article in Journal
Flow Cytometric Methods for the Detection of Intracellular Signaling Proteins and Transcription Factors Reveal Heterogeneity in Differentiating Human B Cell Subsets
Previous Article in Special Issue
Hypoxia Dictates Metabolic Rewiring of Tumors: Implications for Chemoresistance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Fructose Metabolism in Cancer

Department of Bioinformatics and Biochemistry, BRICS, Technische Universität Braunschweig, 38106 Braunschweig, Germany
*
Author to whom correspondence should be addressed.
Cells 2020, 9(12), 2635; https://doi.org/10.3390/cells9122635
Submission received: 6 November 2020 / Revised: 3 December 2020 / Accepted: 4 December 2020 / Published: 8 December 2020

Abstract

:
The interest in fructose metabolism is based on the observation that an increased dietary fructose consumption leads to an increased risk of obesity and metabolic syndrome. In particular, obesity is a known risk factor to develop many types of cancer and there is clinical and experimental evidence that an increased fructose intake promotes cancer growth. The precise mechanism, however, in which fructose induces tumor growth is still not fully understood. In this article, we present an overview of the metabolic pathways that utilize fructose and how fructose metabolism can sustain cancer cell proliferation. Although the degradation of fructose shares many of the enzymes and metabolic intermediates with glucose metabolism through glycolysis, glucose and fructose are metabolized differently. We describe the different metabolic fates of fructose carbons and how they are connected to lipogenesis and nucleotide synthesis. In addition, we discuss how the endogenous production of fructose from glucose via the polyol pathway can be beneficial for cancer cells.

1. Introduction

Fructose occurs naturally in many fruits and vegetables, where it is often bonded to glucose to form the disaccharide sucrose. Hence, it is an integral part of the human diet. With the discovery of the enzymatic conversion of glucose to fructose in 1957 [1], fructose is used as a supplement in the food industry. Fructose is much sweeter than other hexoses, and, therefore, is widely used as a sweetener in processed food and beverages. Since 1970, fructose has been used in the form of high fructose corn syrup (HFCS) leading to an increase of fructose intake in the human diet in the last decades [2]. Fructose consumption, as the combined intake of sucrose and HFCS, increased from 64 g/day in 1970 to 81 g/day in 1997 in the US. The intake of fructose as monosaccharide increased from less than 0.5 g/day to more than 40 g/day [3]. In the US, the sole HFCS intake increased from 0.4 g/day per capita in 1970 to a maximum of 46.6 g/day in 1998, and reduced to 26.8 g/day of HFCS in 2019 [4].
In order to understand the role of fructose in cancer cell metabolism, it is necessary to know and understand how healthy cells metabolize fructose (e.g., the involved metabolic pathways). In the following, we will describe the metabolic fates of fructose in humans and the key enzymes involved in fructose metabolism. Furthermore, we give a short overview of known correlations between fructose consumption and non-cancer diseases like obesity, diabetes, and metabolic syndrome, as they are major risk factors for the development of many different cancer types [5,6], also reviewed in [7]. Subsequently, we will highlight the link between fructose metabolism and cancer cell proliferation, focusing on how fructose is metabolized in different types of cancer, which enzymes are involved, and how fructose metabolism is associated with cancer development and outcome.

2. Physiological Fate(s) of Fructose

The degradation of fructose is usually referred to as fructolysis and is defined as the metabolism of fructose from dietary sources. Although fructolysis shares many of the enzymes and metabolic intermediates with glucose metabolism through glycolysis, glucose and fructose are metabolized differently. While glucose is directly metabolized by virtually every cell type in the human body, fructolysis is assumed to be mainly restricted to the liver. However, small oral loads are metabolized in the small bowel and in small parts extracted by the kidney and other tissues [8]. In the liver, fructose is transported across membranes via passive transport. Here, glucose transporter 2 (GLUT2) is the main fructose transporter. Besides fructose, GLUT2 can also transport glucose and galactose [9]. The sole known fructose specific transporter in the family of glucose transporters is GLUT5, encoded by the gene SLC2A5 and localized mainly in the small intestine [10,11]. While several GLUT transporters are able to transport fructose, GLUT5 is specific for fructose with no ability to transport glucose or galactose [12,13].
The ingested fructose is mainly metabolized via ketohexokinase (KHK), also known as fructokinase in the liver. KHK phosphorylates fructose to fructose-1-phosphate (F1P), which is then hydrolyzed via aldolase B to dihydroxyacetone phosphate (DHAP) and glyceraldehyde (GA) (Figure 1). Both DHAP and GA can be further metabolized via glycolysis in the form of glyeraldehyde-3-phosphate (GA3P). Furthermore, GA can be metabolized to glycerol, which in turn can be phosphorylated to glycerol-3-phosphate (G3P), which can be utilized for lipid synthesis or transporting reducing equivalents into mitochondria via the glycerol-3-phosphate shuttle [14]. Beyond that, fructose can be phosphorylated via hexokinase (HK) to fructose-6-phosphate (F6P). Entering glycolysis, F6P is phosphorylated to fructose-1,6-bisphosphate (F1,6BP) via phosphofructokinase (PFK). Like F1P, F1,6BP is hydrolyzed via aldolase, but this time in DHAP and GA3P (Figure 1). The advantage of fructose degradation via KHK is to bypass PFK. PFK is the rate-limiting enzyme of glycolysis and is inhibited by high levels of ATP and citrate as well as low pH and low oxygen levels. Besides these two pathways with an initial phosphorylation step, fructose can theoretically also be converted into glucose. The so-called polyol pathway consists of two reversible reactions, with sorbitol as an intermediate (Figure 1). Here, the hydrogenation of fructose to sorbitol is catalyzed by sorbitol dehydrogenase (SORD) oxidizing NADH to NAD+. Subsequently, sorbitol is dehydrogenated to glucose via aldo-keto reductase family 1, member B1 (AKR1B1). This reaction needs NADP+ as a cofactor. Since the direction of a metabolic reaction depends on the metabolic levels of its substrates and products [15], the polyol pathway provides the means to produce fructose from glucose endogenously [16], as it is known in diabetic patients.
In summary, for the degradation of fructose, three enzymes seem to be essential: KHK phosphorylating fructose to F1P, HK phosphorylating fructose to F6P, and SORD hydrogenating fructose to sorbitol.

3. Fructose in Disease

Although we will focus on the link between fructose metabolism and cancer, we will briefly mention other diseases affected by fructose metabolism, as some of them are known risk factors for cancer development. For more details, we refer to other reviews [7,17,18,19]. To date, there are three known inborn errors in fructose metabolism: essential or benign fructosuria due to KHK deficiency; hereditary fructose intolerance and fructose-1,6-bisphosphatase (FBPase) deficiency [19].
Essential fructosuria is caused by a loss of function mutation in KHK. This genetic disorder is asymptotic and harmless. Here, ingested fructose is partly excreted in the urine, while the rest is metabolized via hexokinase to F6P in extrahepatic tissue, like adipose tissue and muscle [19,20].
Hereditary fructose intolerance is caused by a deficiency of aldolase B, which converts F1P into DHAP and GA. Loss of aldolase B in combination with a high KHK activity leads to an accumulation of F1P. The high KHK activity depletes the intracellular ATP levels. This lack of ATP leads to increased uric acid production and a release of magnesium, which consequently are a cause for hepatic and renal dysfunction. These toxic effects in hereditary fructose intolerance can be fatal, when sorbitol and fructose are not removed from the diet [19,21,22].
FBPase deficiency is a rare, autosomal recessive disorder affecting one of the key enzymes of gluconeogenesis. Like hereditary fructose intolerance, FBPase deficiency results in F1P accumulation, with severe neurological complications induced by ingesting fructose or glycerol [19,20,23]. Due to the FBPase deficiency, its substrate F1,6BP accumulates, which in turn leads to higher levels of trioses and product inhibition of the aldolase reaction [12,13,20].
In addition to the genetic defects in fructose metabolism, there are several metabolic diseases associated with increased fructose intake. Over the last few years, there were several studies describing the link between fructose consumption and the development of obesity [3,24,25,26,27]. Increased fructose intake leads to an increased energy intake, which in combination with an increased lipid synthesis results in obesity. Beyond that, fructose plays an important role in diabetes and metabolic syndrome, as an increased fructose consumption can lead to insulin resistance [28,29,30].

4. Fructose in Cancer Metabolism

The link between cancer and altered glucose metabolism was already discovered by Otto Warburg around 100 years ago. While aberrant glucose metabolism and the Warburg effect are well investigated [31], the connection between cancer and fructose metabolism needs to be investigated in more detail [7]. Fructose can be metabolized in many different ways (Figure 1). Some of these metabolic pathways overlap with those of glucose degradation and play a central role in cell growth and survival. Therefore, it is not surprising that fructose, like glucose, can affect the growth, proliferation, and survival of cancer cells [32,33]. Before we discuss important metabolic pathways of fructose metabolism in more detail, we will shortly highlight recent studies connecting the fructose transporter GLUT5 to cancer development and progression. In patients with lung adenocarcinoma, a poor prognosis is correlated with an upregulation of GLUT5 [34]. The authors demonstrate that the depletion of GLUT5 attenuated cell proliferation and invasion, and increased apoptosis. The overexpression of GLUT5, on the other hand, enhanced cell proliferation, migration, invasion, and tumorigenesis [34]. Enhanced fructose utilization mediated by GLUT5 was also found in acute myeloid leukemia. The increased fructose uptake resulted in an increased proliferation rate, colony growth, and enhanced migration and invasion. Consequently, high expression of GLUT5 and enhanced fructose utilization are associated with poor outcomes and exacerbate leukemic phenotypes [11]. In addition, GLUT5 is overexpressed in several types of cancer, e.g., glioblastoma, colon, liver, lung, breast, and prostate [35,36].

4.1. Polyol Pathway

The polyol pathway describes the two-step conversion of glucose to fructose [37,38,39]. First, glucose is reduced to sorbitol with the use of NADPH, then sorbitol is oxidized to fructose producing NADH from NAD+. The polyol pathway is well known for its pathological implications in diabetes [40,41,42,43]. Here, in hyperglycemic state, sorbitol sythesis increases, thereby consuming most of the NADPH, which in turn results in increased oxidative stress [40,43,44,45]. The polyol pathway is interesting for fructose metabolism because it provides the means to either convert fructose to glucose or to endogenously produce fructose from glucose. Both of these options provide an interesting perspective on cancer cell metabolism.
Converting fructose to glucose via the polyol pathway can theoretically directly fuel glycolysis. This would enable fructolysis detached from KHK. In humans, the conversion of fructose via KHK is assumed to be mainly restricted to liver [46,47]. However, in tumors, KHK is apparently expressed more frequently [48,49,50]. To our knowledge, there is no experimental evidence that this pathway is physiologically relevant. Converting fructose to glucose via the polyol pathway would require low glucose and low NADPH levels. The reverse flux, from glucose to fructose, enabling endogenous fructose production, seems to be more physiologically relevant [51,52]. It is important to note that the endogenous fructose production with subsequent KHK activity bypasses the rate-limiting step in glycolysis, catalyzed by PFK, en route to GA3P production. The phosphorylation of F6P to F1,6BP mediated by PFK is inhibited by high levels of ATP or citrate and a low pH (Figure 1). This bypassing effect was recently observed in naked mole-rats by Park et al. [53]. In their study, the authors observed that naked mole-rats can survive a prolonged time under low oxygen levels and even anoxia because they are able to utilize fructose to fuel glycolysis for ATP generation. Low oxygen levels usually lead to an inhibition of oxidative phosphorylation and an increase in glycolytic activity and subsequently to a lower pH, which in turn would inhibit PFK. The authors concluded that fueling glycolysis with fructose via KHK avoided the PFK feedback inhibition by low pH, as GA and DHAP enter glycolysis downstream of PFK (Figure 1). Expanding on that concept, it is not too far fetched that cancer cells would employ such a mechanism. Most solid tumors have an inadequate supply of oxygen, resulting in hypoxic and an acidic environment [54]. In that context, polyol pathway activity could provide the means for cancer cells to produce fructose from glucose, bypassing PFK and increasing the glycolytic rate.
A direct link between polyol pathway activity and cancer was discovered by Schwab et. al [55]. In their study, the authors showed a strong correlation between AKR1B1 expression and epithelial-to-mesenchymal transition (EMT) in lung cancer patients and in an EMT-driven colon cancer mouse model. In in vitro experiments, they detected increased AKR1B1 levels in mesenchymal-like cancer cells, while AKR1B1 knockdown was sufficient to revert EMT. They observed similar results in EMT suppression when they targeted SORD. Furthermore, Schwab and colleagues showed that glucose-induced activation of the polyol pathway controls EMT via TGF β autocrine stimulation. TGF β levels in AKR1B1 and SORD knockdown cells were reduced and EMT markers could be rescued adding exogenous TGF β [55]. In line with these results, there are a couple more studies pointing towards a correlation between the polyol pathway, in particular AKR1B1, and cancer. AKR1B1 is overexpressed in breast, ovarian, cervical, and rectal cancer [56]. In colorectal cancer cells, AKR1B1 affects cellular proliferation and cell cycle progression. In addition, AKR1B1 enhances cell motility and NF- κ B activity, resulting in a poor prognosis in colorectal cancer patients [57].
While there is mounting experimental evidence showing the connection of aldo reductases and cancer, little is known about the connection between SORD and cancer. In addition to the findings by Schwab and coworkers showing the association between SORD expression and EMT [55], Uzozie and colleagues analyzed SORD overexpression in precancerous colorectal neoplasms [58]. Here, increased SORD activity in adenomas and cancer cell lines lead to changes in expression of AKR1B1 and KHK. When SORD is upregulated, AKR1B1 was decreased and KHK was upregulated.
In summary, endogenous fructose production via the polyol pathway seems to be an important part of cancer cell metabolism. The underlying molecular mechanism, however, remains elusive. Besides bypassing PFK regulation and entering glycolysis for ATP production, endogenously produced fructose could enter several other pathways important for cancer cell physiology such as the hexosamine pathway [59], the pentose phosphate pathway [60], or de novo lipogenesis [61].

4.2. The Pentose Phosphate Pathway

In rapidly dividing cells, such as cancer cells, it has been shown that the pentose phosphate pathway (PPP) is critical to sustain high proliferation rates [62] and that fructose metabolism plays an important role in it [63]. Generally, the PPP has two important cellular functions: (1) to provide metabolic precursors for macromolecule synthesis (ribose 5-phosphate for nucleotide and erythrose 4-phosphate for amino acid biosynthesis) and (2) to provide NADPH for anabolism and redox homeostasis. It can be broken down into two branches: the oxidative and the non-oxidative branch (Figure 2). While the oxidative branch has three irreversible reactions, the non-oxidative branch comprises a series of reversible reactions that connect the PPP to glycolysis.
In cancer cells, most of the pentose phosphates, used for de novo nucleotide synthesis, are derived from the PPP [62]. As described above (Figure 1), fructose can be metabolized in various ways, and, therefore, can potentially enter the oxidative as well as the non-oxidative branch of the PPP. However, it has been shown that fructose is preferentially metabolized via the non-oxidative branch in pancreatic cancer cells [63]. As such, the fructose carbon backbone can enter the non-oxidative branch as glyceraldehyd-3-phosphate or fructose-6-phosphate, catalyzed by the transketolase (TKT) enzyme (Figure 2). The authors showed that, in fructose treated cells, TKT protein expression increased 200% compared to glucose treated cells. In line with that observation, the authors observed that 13 C labeled fructose was metabolized at a 250% higher rate than glucose via TKT to synthesize nucleic acids [63]. Using the non-oxidative branch of the PPP for ribose synthesis has the advantage that it is uncoupled from the generation of NADPH. This is useful when the metabolic requirement of ribose synthesis exceeds that of NADPH. Moreover, the non-oxidative branch has a neutral carbon balance as it is not producing CO2 like the oxidative branch (Figure 2).
The upregulation of TKT, and thus acceleration of the non-oxidative PPP, seems to be a general feature of cancer cells [64,65], even with glucose as the carbon source. Based on 13 C-tracer experiments, it was shown in vitro that around 80% of ribonucleotides are derived from the non-oxidative PPP [66]. In chronic myeloid leukemia, it was shown that a HIF-1 α dependent induction of TKT is associated with imatinib-resistance [67]. Inhibition of TKT activity with oxythiamine enhanced imatinib sensitivity of tumor cells [67]. In addition, a homologue of TKL, TKL-like protein (TKLL1) was detected in tumor tissue and its expression level has been correlated with cancer progression [68,69]. Although it was shown that TKLL1 possesses TKT activity, it is debated whether TKLL1 participates in the PPP [70].
It is worth underlining the importance of the non-oxidative branch for cancer cells, and it was shown that a G6PD deficiency does not reduce the risk of getting cancer [71], showing that NADPH requirements can be fulfilled by other pathways (e.g., serine-driven one-carbon metabolism [72]) and that the non-oxidative branch is sufficient to sustain nucleotide synthesis. As stated above, it is assumed that most of the ribonucleotides are derived from the non-oxidative branch and that fructose is preferentially metabolized via the non-oxidative branch of the PPP. This raises the possibility that cancer cells use glucose for endogenous fructose production via the polyol pathway and then subsequently use the fructose to fuel the non-oxidative PPP for nucleic acid synthesis.

4.3. De Novo Lipogenesis

Besides an increased demand of ribonucleotides, proliferating cancer cells require a constant supply of fatty acids for the biogenesis of more complex lipids (e.g., membrane lipids) [73]. Cancer cells can either increase their uptake of lipids or increase the metabolic flux into de novo lipogenesis (DNL) [74]. Generally, DNL refers to the metabolic pathway that synthesizes fatty acids from excess carbohydrates, which are then subsequently incorporated into triglycerides for energy storage and later oxidation. In humans, DNL mainly takes place in liver and adipose tissue and is mediated by the cytosolic enzymes acetyl-CoA carboxylase (ACC) and fatty acid synthase (FASN) [61]. In many cancer types, FASN is overexpressed [75], and thus the majority of fatty acids in these cells are assumed to be derived from DNL [76,77,78].
Fatty acid synthesis starts with the initialization of a new acyl chain by FASN from acetyl-CoA which is then subsequently elongated using malonyl-CoA generated by ACC (Figure 3). In general, the acetyl-CoA is derived from citrate by ATP-citrate lyase (ACLY), or from acetate by acyl-coenzyme A synthetase short-chain family member 2 (ACSS2). The main product of DNL is the saturated 16 carbon fatty acid palmitate, which can be used further for the synthesis of more complex fatty acids or triglycerides as well as phospholipids by esterification with glycerol.
The interest in fructose as a driver of DNL is based on the observation that a higher dietary intake of fructose leads to an increased incidence of non-alcoholic fatty liver disease [79]. In this context, Silva and colleagues showed that, in the liver of mice, fructose contributed significantly more to saturated fatty acid synthesis compared to glucose. About 30% of fructose is used for hepatic de novo fatty acid synthesis in the form of acetyl-CoA, while the contribution of glucose is significantly lower (~10%) [80]. While this paper elegantly showed the increased fructose carbon contribution to DNL, it does not show the metabolic pathway used to convert the fructose carbon into acetyl-CoA. As shown in Figure 3, the fructose carbon backbone can enter glycolysis and can eventually be metabolized to acetyl-CoA. Recently, Zhao and colleagues, however, showed that the microbiome contributes to the supply of lipogenic acetyl-CoA by converting fructose to acetate [81]. Using in vivo stable isotope analysis, they observed that in liver-specific Acly knockout mice, fructose carbons were still incorporated into fatty acids and that fructose carbon contribution to hepatic acetyl-CoA was significantly reduced when the microbiome was depleted with an antibiotic cocktail. The authors concluded that dietary fructose is converted to acetate by the gut microbiome, making DNL from fructose independent of Acly activity.
Interestingly, acetyl-CoA is not the sole possibility of how fructose carbons can be utilized for lipid synthesis. Fructose carbons can end up downstream in glycerol (Figure 3), providing the backbone for triglycerides and phospholipids. Compared to glucose, it was shown that fructose provides a greater amount of carbon for G3P synthesis in liver [80]. The authors showed in 13C fructose fed mice that, in the liver, about 60% of newly synthesized glycerol is derived from fructose [80]. Not only is fructose preferentially metabolized to triglycerides, but it also increases triglyceride levels [82]. Triglyceride levels are elevated in several types of cancer [83,84] or associated with an increased risk of cancer [85]. These findings demonstrate that fructose is used as a carbon source for fatty acid and triglyceride synthesis and therefore provides components to sustain cancer cell growth and proliferation. However, how much oral fructose consumption actually contributes to triglyceride synthesis in healthy humans is still debated. It was estimated to contribute <1% to plasma triglycerides, but increased DNL in general as assessed by 13 C labelled acetate studies [86], suggesting that fructose may drive DNL without donating carbons.
Over the last few years, several research groups have demonstrated that fructose consumption increases DNL [87] and expression levels of the DNL enzymes, ACC [88], FASN [89], and stearoyl CoA desaturase-1 (SCD1) [90]. In fructose-fed rats, it was observed that the expression levels of FASN and SCD1 were significantly increased compared to controls [91]. Similar effects were observed for other lipogenic enzymes [78]. Recently, it was observed that fructose, even at a moderate dose, induces fatty acid synthesis and can enhance tumorigenesis in mice [92]. All this suggests that fructose not only provides carbon for the formation of fatty acids, but also has a signaling function in DNL. However, little is known about the mechanistic role of fructose on DNL enzyme expression. A prominent regulator of DNL is sterol regulatory element-binding protein 1c (SREBF1), which induces many signals for DNL in response to carbohydrate intake [93]. In addition, carbohydrate regulatory element-binding protein (ChREBP) is activated by carbohydrate intake and controls the expression of several lipogenic enzymes [94]. In rat liver, it was observed that nucleus SREBF1 and ChREBP activity increased when dietary glucose is replaced by fructose [95]. The authors showed increased ChREBP binding to ChoRE, a promotor of lipogenic enzymes, and increased active SREBF1 [95]. It has been suggested that hexose-phosphate metabolites are important for ChREBP activation [96,97], highlighting the importance of KHK in the fructose dependent activation of ChREBP. In liver-specific Acly knockout mice consuming a mixture of fructose:glucose, it was observed that ChREBP and DNL gene expression regulation was independent of acetyl-CoA metabolism [81]. The results are in line with experiments showing that KHK knockout mice are protected from fructose-induced fatty liver [98,99].
A recent study shows both the carbon contribution as well as signaling effect of fructose in hepatic DNL [100]. In transgenic mice, a high-fructose diet led to higher cytosolic acetyl-CoA pool, increased liver triglycerides, and increased expression of SREBF1, ChREBP, ACC1, and FASN. The authors explained the signaling effects based on the interplay between intestinal epithelial cells, macrophages, and hepatocytes. In intestinal epithelial cells, a high-fructose diet resulted in high F1P levels and subsequently in deregulated protein N-glycosylation, triggering ER stress and intestinal inflammation. Inflammatory signals activated MyD88 signaling in macrophages and subsequent TNF release. TNF then induced SREBF1 signaling in hepatocytes [100].
In summary, these findings suggest that, in cancer cells, fructose can play an important role as a carbon precursor for fatty acid and triglyceride synthesis as well as in the functional regulation of DNL and its gene expression (Figure 4).
In addition, DNL plays also an important role in cancer cell survival. In breast cancer cells, inhibition of ACC and FASN results in impaired cancer cell migration and invasion [101]. Furthermore, the authors showed a strong inhibition of cancer cell proliferation in breast cancer cells and hepatocellular carcinoma cells when ACC and FASN were inhibited. In addition, DNL promotes membrane lipid saturation and therefore protects cancer cells from free radicals and chemotherapeutics [102]. These findings make fructose metabolism, especially DNL, an interesting target for anti-cancer therapy and suggest that lipid synthesis play an important role in cancer cell survival and biology.

5. Outlook

Compared to glucose, the connection of fructose to cancer cell metabolism has long been overlooked. However, fructose can be used to generate ATP in glycolysis, donate carbon for nucleotide and lipid synthesis, and can provide signaling cues to sustain cancer cell proliferation. Indeed, in recent years, there has been mounting experimental as well as clinical evidence that increased fructose consumption can support tumorigenesis. In addition, several research groups have shown that the endogenous fructose production from glucose via the polyol pathway plays an important role in the formation of metastasis. Looking ahead, it will be interesting to quantify the metabolic flux from glucose to fructose and to subsequently follow the fate of these carbons in central metabolism. This will prove challenging, as standard metabolic flux techniques (e.g., stable-isotope assisted metabolomics) will have difficulties to distinguish the polyol pathway flux from the normal glycolytic flux. These methods need to be combined with appropriate knockout or overexpression models.
Another interesting question is whether fructose metabolism can be a suitable drug target for the treatment of cancer. Here, KHK could become an interesting therapeutic target. Many cancer cells overexpress KHK, and essential fructosuria, the genetic disorder caused by a loss of function mutation in KHK, is clinically asymptomatic and harmless, further supporting the notion that clinical inhibition of KHK may be tolerated in cancer patients.

Author Contributions

Conceptualization, N.K. and A.W.; writing—original draft preparation, N.K. and A.W.; visualization, N.K. and A.W.; writing—review and editing N.K. and A.W.; supervision A.W.; funding acquisition, A.W. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge funding by the Ministry of Science and Culture (MWK) of Lower Saxony: (1) SMART BIOTECS alliance between the Technische Universität Braunschweig and the Leibnitz Universität Hannover and (2) Research cooperation Lower Saxony-Israel; the Federal Ministry of Education and Research: PeriNAA-01ZX1916B, and the German Research Foundation and the Open Access Publication Funds of Technische Universität Braunschweig.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
ACCacetyl-CoA carboxylase
ACLYATP citrate lyase
ACSS2acyl-coenzyme A synthetase short-chain family member 2
AKG α -ketoglutarate
AKR1B1Aldo-Keto Reductase Family 1, Member B1
ALDOAldolase
ChREBPCarbohydrate-Responsive Element-Binding Protein
ChoRECarbohydrate-Responsive Element
DHAPDihydroxyacetone Phosphate
DNLDe Novo Lipogenesis
E4PErythrose 4-Phosphate
EMTEpithelial-to-Mesenchymal Transition
F1PFructose-1-Phosphate
F6PFructose-6-Phosphate
FASNFatty Acid Synthase
F1,6BPFructose-1,6-Bisphosphate
FBPaseFructose-1,6-Bisphosphatase
G3PGlycerol-3-Phosphate
GAGlyceraldehyde
GA3PGlyceraldehyde-3-Phosphate
GLUTGlucose Transporter
G6PDGlucose-6-Phosphate Dehydrogenase
GPDHGlycerol 3-Phosphate Dehydrogenase
GPIGlucose-6-Phosphate Isomerase
GSHGlutathione
GSSGGlutathione disulfide
HFCSHigh Fructose Corn Syrup
HIF1- α Hypoxia-inducible factor 1-alpha
HKHexokinase
KHKKetohexokinase
NF- κ BNuclear Factor kappa-light-chain-enhancer of activated B-cells
6PGL6-phosphogluconolactonase
6PGDH6-phosphogluconate dehydrogenase
PPPPentose Phosphate pathway
PFKPhosphofructokinase
RPERibulose 5-Phosphate 3-Epimerase
RPIRibose-5-Phosphate Isomerase
SCD1Stearoyl-CoA desaturase
SORDSorbitol Dehydrogenase
SRESterol Regulatory Element
SREBFSterol Regulatory Element-Binding protein
TAGTriacylglyceride
TALDOTransaldolase
TCATricarboxylic Acid Cycle
TGF β Transforming Growth Factor β
TKTriose Kinase
TKLL1Transketolase-Like Protein
TKTTransketolase
TNFTumor Necrosis Factor
TPITriose-Phosphate Isomerase
X5PXylulose 5-Phosphate

References

  1. Marshall, R.O.; Kooi, E.R.; Moffett, G.M. Enzymatic Conversion of d-Glucose to d-Fructose. Science 1957, 125, 648–649. [Google Scholar] [CrossRef] [PubMed]
  2. White, J.S. Straight talk about high-fructose corn syrup: What it is and what it ain’t. Am. J. Clin. Nutr. 2008, 88, 1716S–1721S. [Google Scholar] [CrossRef] [PubMed]
  3. Elliott, S.S.; Keim, N.L.; Stern, J.S.; Teff, K.; Havel, P.J. Fructose, weight gain, and the insulin resistance syndrome. Am. J. Clin. Nutr. 2002, 76, 911–922. [Google Scholar] [CrossRef] [PubMed]
  4. Sowell, A. Table 52-High-fructose corn syrup: Estimated number of per capita calories consumed daily, by calendar year. In Sugar and Sweeteners Yearbook Tables; United States Department of Agriculture: Washington, DC, USA, 2020. [Google Scholar]
  5. Gallagher, E.J.; LeRoith, D. Obesity and Diabetes: The Increased Risk of Cancer and Cancer-Related Mortality. Physiol. Rev. 2015, 95, 727–748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Uzunlulu, M.; Telci Caklili, O.; Oguz, A. Association between Metabolic Syndrome and Cancer. Ann. Nutr. Metab. 2016, 68, 173–179. [Google Scholar] [CrossRef]
  7. Nakagawa, T.; Lanaspa, M.A.; Millan, I.S.; Fini, M.; Rivard, C.J.; Sanchez-Lozada, L.G.; Andres-Hernando, A.; Tolan, D.R.; Johnson, R.J. Fructose contributes to the Warburg effect for cancer growth. Cancer Metab. 2020, 8. [Google Scholar] [CrossRef]
  8. Tappy, L.; Rosset, R. Health outcomes of a high fructose intake: The importance of physical activity. J. Physiol. 2019, 597, 3561–3571. [Google Scholar] [CrossRef] [Green Version]
  9. Manolescu, A.R.; Witkowska, K.; Kinnaird, A.; Cessford, T.; Cheeseman, C. Facilitated hexose transporters: New perspectives on form and function. Physiology 2007, 22, 234–240. [Google Scholar] [CrossRef]
  10. Douard, V.; Ferraris, R.P. Regulation of the fructose transporter GLUT5 in health and disease. Am. J. Physiol. Endocrinol. Metab. 2008, 295, E227–E237. [Google Scholar] [CrossRef] [Green Version]
  11. Chen, W.L.; Wang, Y.Y.; Zhao, A.; Xia, L.; Xie, G.; Su, M.; Zhao, L.; Liu, J.; Qu, C.; Wei, R.; et al. Enhanced Fructose Utilization Mediated by SLC2A5 Is a Unique Metabolic Feature of Acute Myeloid Leukemia with Therapeutic Potential. Cancer Cell 2016, 30, 779–791. [Google Scholar] [CrossRef] [Green Version]
  12. Thorens, B.; Mueckler, M. Glucose transporters in the 21st Century. Am. J. Physiol. Endocrinol. Metab. 2010, 298, E141–E145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Douard, V.; Ferraris, R.P. The role of fructose transporters in diseases linked to excessive fructose intake. J. Physiol. 2012, 591, 401–414. [Google Scholar] [CrossRef] [PubMed]
  14. Mráček, T.; Drahota, Z.; Houštěk, J. The function and the role of the mitochondrial glycerol-3-phosphate dehydrogenase in mammalian tissues. Biochim. Biophys. Acta (Bba) Bioenergy 2013, 1827, 401–410. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Wegner, A.; Meiser, J.; Weindl, D.; Hiller, K. How metabolites modulate metabolic flux. Curr. Opin. Biotechnol. 2015, 34, 16–22. [Google Scholar] [CrossRef] [Green Version]
  16. Lanaspa, M.A.; Ishimoto, T.; Cicerchi, C.; Tamura, Y.; Roncal-Jimenez, C.A.; Chen, W.; Tanabe, K.; Andres-Hernando, A.; Orlicky, D.J.; Finol, E.; et al. Endogenous fructose production and fructokinase activation mediate renal injury in diabetic nephropathy. J. Am. Soc. Nephrol. JASN 2014, 25, 2526–2538. [Google Scholar] [CrossRef] [Green Version]
  17. Van den Berghe, G. Inborn errors of fructose metabolism. Annu. Rev. Nutr. 1994, 14, 41–58. [Google Scholar] [CrossRef]
  18. Steinmann, B.; Santer, R. Disorders of Fructose Metabolism. In Inborn Metabolic Diseases; Springer: Berlin/Heidelberg, Germany, 2012; pp. 157–165. [Google Scholar] [CrossRef]
  19. Tran, C. Inborn Errors of Fructose Metabolism. What Can We Learn from Them? Nutrients 2017, 9, 356. [Google Scholar] [CrossRef]
  20. Steinmann, B.; Santer, R. Disorders of Fructose Metabolism. In Inborn Metabolic Diseases; Saudubray, J.M., Baumgartner, M., Walter, J., Eds.; Springer: Berlin/Heidelberg, Germany, 2016; pp. 161–168. [Google Scholar] [CrossRef]
  21. Gitzelmann, R.; Baerlocher, K.; Prader, A. Hereditary defects of fructose and galactose metabolism. Monatsschrift Fur Kinderheilkd. 1973, 121, 174–180. [Google Scholar]
  22. Froesch, E.R. Disorders of fructose metabolism. Clin. Endocrinol. Metab. 1976, 5, 599–611. [Google Scholar] [CrossRef] [Green Version]
  23. Hasegawa, Y.; Kikawa, Y.; Miyamaoto, J.; Sugimoto, S.; Adachi, M.; Ohura, T.; Mayumi, M. Intravenous glycerol therapy should not be used in patients with unrecognized fructose-1,6-bisphosphatase deficiency. Pediatr. Int. 2003, 45, 5–9. [Google Scholar] [CrossRef]
  24. Lakhan, S.E.; Kirchgessner, A. The emerging role of dietary fructose in obesity and cognitive decline. Nutr. J. 2013, 12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Livesey, G. Fructose, Obesity, and Related Epidemiology. Crit. Rev. Food Sci. Nutr. 2010, 50, 26–28. [Google Scholar] [CrossRef]
  26. Pereira, R.; Botezelli, J.; da Cruz Rodrigues, K.; Mekary, R.; Cintra, D.; Pauli, J.; da Silva, A.; Ropelle, E.; de Moura, L. Fructose Consumption in the Development of Obesity and the Effects of Different Protocols of Physical Exercise on the Hepatic Metabolism. Nutrients 2017, 9, 405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Tappy, L. Fructose-containing caloric sweeteners as a cause of obesity and metabolic disorders. J. Exp. Biol. 2018, 221, jeb164202. [Google Scholar] [CrossRef] [Green Version]
  28. Bantle, J.P. Dietary Fructose and Metabolic Syndrome and Diabetes. J. Nutr. 2009, 139, 1263S–1268S. [Google Scholar] [CrossRef] [Green Version]
  29. Gerrits, P.M.; Tsalikian, E. Diabetes and fructose metabolism. Am. J. Clin. Nutr. 1993, 58, 796S–799S. [Google Scholar] [CrossRef]
  30. Kawasaki, T.; Akanuma, H.; Yamanouchi, T. Increased Fructose Concentrations in Blood and Urine in Patients With Diabetes. Diabetes Care 2002, 25, 353–357. [Google Scholar] [CrossRef] [Green Version]
  31. Vander Heiden, M.G.; Cantley, L.C.; Thompson, C.B. Understanding the Warburg effect: The metabolic requirements of cell proliferation. Science 2009, 324, 1029–1033. [Google Scholar] [CrossRef] [Green Version]
  32. Liu, H.; Heaney, A.P. Refined fructose and cancer. Expert Opin. Ther. Targets 2011, 15, 1049–1059. [Google Scholar] [CrossRef]
  33. Boroughs, L.K.; DeBerardinis, R.J. Metabolic pathways promoting cancer cell survival and growth. Nat. Cell Biol. 2015, 17, 351–359. [Google Scholar] [CrossRef] [Green Version]
  34. Weng, Y.; Fan, X.; Bai, Y.; Wang, S.; Huang, H.; Yang, H.; Zhu, J.; Zhang, F. SLC2A5 promotes lung adenocarcinoma cell growth and metastasis by enhancing fructose utilization. Cell Death Discov. 2018, 4. [Google Scholar] [CrossRef] [PubMed]
  35. Godoy, A.; Ulloa, V.; Rodríguez, F.; Reinicke, K.; Yañez, A.J.; de los Angeles García, M.; Medina, R.A.; Carrasco, M.; Barberis, S.; Castro, T.; 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, 614–627. [Google Scholar] [CrossRef]
  36. Zamora-Leon, S.P.; Golde, D.W.; Concha, I.I.; Rivas, C.I.; Delgado-Lopez, F.; Baselga, J.; Nualart, F.; Vera, J.C. Expression of the fructose transporter GLUT5 in human breast cancer. Proc. Natl. Acad. Sci. USA 1996, 93, 1847–1852. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. HEYNINGEN, R.V. Formation of Polyols by the Lens of the Rat with ‘Sugar’ Cataract. Nature 1959, 184, 194–195. [Google Scholar] [CrossRef]
  38. Kador, P.F.; Kinoshita, J.H. Role of aldose reductase in the development of diabetes-associated complications. Am. J. Med. 1985, 79, 8–12. [Google Scholar] [CrossRef]
  39. Kinoshita, J.H. A thirty year journey in the polyol pathway. Exp. Eye Res. 1990, 50, 567–573. [Google Scholar] [CrossRef]
  40. LEE, A.Y.W.; CHUNG, S.S.M. Contributions of polyol pathway to oxidative stress in diabetic cataract. FASEB J. 1999, 13, 23–30. [Google Scholar] [CrossRef] [Green Version]
  41. Dunlop, M. Aldose reductase and the role of the polyol pathway in diabetic nephropathy. Kidney Int. 2000, 58, S3–S12. [Google Scholar] [CrossRef] [Green Version]
  42. Oates, P.J. Polyol pathway and diabetic peripheral neuropathy. In International Review of Neurobiology; Elsevier: Amsterdam, The Netherlands, 2002; pp. 325–392. [Google Scholar] [CrossRef]
  43. Chung, S.S. Contribution of Polyol Pathway to Diabetes-Induced Oxidative Stress. J. Am. Soc. Nephrol. 2003, 14, 233–236. [Google Scholar] [CrossRef] [Green Version]
  44. Obrosova, I.G. Increased Sorbitol Pathway Activity Generates Oxidative Stress in Tissue Sites for Diabetic Complications. Antioxid. Redox Signal. 2005, 7, 1543–1552. [Google Scholar] [CrossRef]
  45. Yan, L.-J. Redox imbalance stress in diabetes mellitus: Role of the polyol pathway. Anim. Model. Exp. Med. 2018, 1, 7–13. [Google Scholar] [CrossRef] [PubMed]
  46. Hayward, B.E.; Bonthron, D.T. Structure and alternative splicing of the ketohexokinase gene. Eur. J. Biochem. 1998, 257, 85–91. [Google Scholar] [CrossRef] [PubMed]
  47. Diggle, C.P.; Shires, M.; Leitch, D.; Brooke, D.; Carr, I.M.; Markham, A.F.; Hayward, B.E.; Asipu, A.; Bonthron, D.T. Ketohexokinase: Expression and Localization of the Principal Fructose-metabolizing Enzyme. J. Histochem. Cytochem. 2009, 57, 763–774. [Google Scholar] [CrossRef] [Green Version]
  48. 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, 1298–1306. [Google Scholar] [CrossRef] [PubMed]
  49. Yang, X.; Shao, F.; Shi, S.; Feng, X.; Wang, W.; Wang, Y.; Guo, W.; Wang, J.; Gao, S.; Gao, 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. [Google Scholar] [CrossRef] [Green Version]
  50. Kim, J.; Kang, J.; Kang, Y.L.; Woo, J.; Kim, Y.; Huh, J.; Park, J.W. Ketohexokinase-A acts as a nuclear protein kinase that mediates fructose-induced metastasis in breast cancer. Nat. Commun. 2020, 11. [Google Scholar] [CrossRef]
  51. Lanaspa, M.A.; Ishimoto, T.; Li, N.; Cicerchi, C.; Orlicky, D.J.; Ruzycki, P.; Rivard, C.; Inaba, S.; Roncal-Jimenez, C.A.; Bales, E.S.; et al. Endogenous fructose production and metabolism in the liver contributes to the development of metabolic syndrome. Nat. Commun. 2013, 4. [Google Scholar] [CrossRef]
  52. Andres-Hernando, A.; Johnson, R.J.; Lanaspa, M.A. Endogenous fructose production. Curr. Opin. Clin. Nutr. Metab. Care 2019, 22, 289–294. [Google Scholar] [CrossRef]
  53. Park, T.J.; Reznick, J.; Peterson, B.L.; Blass, G.; Omerbašić, D.; Bennett, N.C.; Kuich, P.H.J.L.; Zasada, C.; Browe, B.M.; Hamann, W.; et al. Fructose-driven glycolysis supports anoxia resistance in the naked mole-rat. Science 2017, 356, 307–311. [Google Scholar] [CrossRef] [Green Version]
  54. Höckel, M.; Vaupel, P. Tumor hypoxia: Definitions and current clinical, biologic, and molecular aspects. J. Natl. Cancer Inst. 2001, 93, 266–276. [Google Scholar] [CrossRef] [Green Version]
  55. Schwab, A.; Siddiqui, A.; Vazakidou, M.E.; Napoli, F.; Böttcher, M.; Menchicchi, B.; Raza, U.; Saatci, Ö.; Krebs, A.M.; Ferrazzi, F.; et al. Polyol Pathway Links Glucose Metabolism to the Aggressiveness of Cancer Cells. Cancer Res. 2018, 78, 1604–1618. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Saraswat, M.; Mrudula, T.; Kumar, P.U.; Suneetha, A.; Rao Rao, T.S.; Srinivasulu, M.; Reddy, B. Overexpression of aldose reductase in human cancer tissues. Med. Sci. Monit. Int. Med. J. Exp. Clin. Res. 2006, 12, CR525–CR529. [Google Scholar]
  57. Taskoparan, B.; Seza, E.G.; Demirkol, S.; Tuncer, S.; Stefek, M.; Gure, A.O.; Banerjee, S. Opposing roles of the aldo-keto reductases AKR1B1 and AKR1B10 in colorectal cancer. Cell. Oncol. 2017, 40, 563–578. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Uzozie, A.; Nanni, P.; Staiano, T.; Grossmann, J.; Barkow-Oesterreicher, S.; Shay, J.W.; Tiwari, A.; Buffoli, F.; Laczko, E.; Marra, G. Sorbitol Dehydrogenase Overexpression and Other Aspects of Dysregulated Protein Expression in Human Precancerous Colorectal Neoplasms: A Quantitative Proteomics Study. Mol. Cell. Proteom. 2014, 13, 1198–1218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Chiaradonna, F.; Ricciardiello, F.; Palorini, R. The Nutrient-Sensing Hexosamine Biosynthetic Pathway as the Hub of Cancer Metabolic Rewiring. Cells 2018, 7, 53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Stincone, A.; Prigione, A.; Cramer, T.; Wamelink, M.M.C.; Campbell, K.; Cheung, E.; Olin-Sandoval, V.; Grüning, N.M.; Krüger, A.; Tauqeer Alam, M.; et al. The return of metabolism: Biochemistry and physiology of the pentose phosphate pathway. Biol. Rev. Camb. Philos. Soc. 2015, 90, 927–963. [Google Scholar] [CrossRef] [Green Version]
  61. Ameer, F.; Scandiuzzi, L.; Hasnain, S.; Kalbacher, H.; Zaidi, N. De novo lipogenesis in health and disease. Metabolism 2014, 63, 895–902. [Google Scholar] [CrossRef]
  62. Raïs, B.; Comin, B.; Puigjaner, J.; Brandes, J.L.; Creppy, E.; Saboureau, D.; Ennamany, R.; Lee, W.N.P.; Boros, L.G.; Cascante, M. Oxythiamine and dehydroepiandrosterone induce a G1 phase cycle arrest in Ehrlich’s tumor cells through inhibition of the pentose cycle. FEBS Lett. 1999, 456, 113–118. [Google Scholar] [CrossRef] [Green Version]
  63. Liu, H.; Huang, D.; McArthur, D.L.; Boros, L.G.; Nissen, N.; Heaney, A.P. Fructose Induces Transketolase Flux to Promote Pancreatic Cancer Growth. Cancer Res. 2010, 70, 6368–6376. [Google Scholar] [CrossRef] [Green Version]
  64. Ricciardelli, C.; Lokman, N.A.; Cheruvu, S.; Tan, I.A.; Ween, M.P.; Pyragius, C.E.; Ruszkiewicz, A.; Hoffmann, P.; Oehler, M.K. Transketolase is upregulated in metastatic peritoneal implants and promotes ovarian cancer cell proliferation. Clin. Exp. Metastasis 2015, 32, 441–455. [Google Scholar] [CrossRef]
  65. Qin, Z.; Xiang, C.; Zhong, F.; Liu, Y.; Dong, Q.; Li, K.; Shi, W.; Ding, C.; Qin, L.; He, F. Transketolase (TKT) activity and nuclear localization promote hepatocellular carcinoma in a metabolic and a non-metabolic manner. J. Exp. Clin. Cancer Res. 2019, 38. [Google Scholar] [CrossRef] [PubMed]
  66. Boros, L.G.; Puigjaner, J.; Cascante, M.; Lee, W.N.; Brandes, J.L.; Bassilian, S.; Yusuf, F.I.; Williams, R.D.; Muscarella, P.; Melvin, W.S.; et al. Oxythiamine and dehydroepiandrosterone inhibit the nonoxidative synthesis of ribose and tumor cell proliferation. Cancer Res. 1997, 57, 4242–4248. [Google Scholar] [PubMed]
  67. Zhao, F.; Mancuso, A.; Bui, T.V.; Tong, X.; Gruber, J.J.; Swider, C.R.; Sanchez, P.V.; Lum, J.J.; Sayed, N.; Melo, J.V.; et al. Imatinib resistance associated with BCR-ABL upregulation is dependent on HIF-1α-induced metabolic reprograming. Oncogene 2010, 29, 2962–2972. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Diaz-Moralli, S.; Tarrado-Castellarnau, M.; Alenda, C.; Castells, A.; Cascante, M. Transketolase-Like 1 Expression Is Modulated during Colorectal Cancer Progression and Metastasis Formation. PLoS ONE 2011, 6, e25323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Kayser, G.; Sienel, W.; Kubitz, B.; Mattern, D.; Stickeler, E.; Passlick, B.; Werner, M.; zur Hausen, A. Poor outcome in primary non-small cell lung cancers is predicted by transketolase TKTL1 expression. Pathology 2011, 43, 719–724. [Google Scholar] [CrossRef]
  70. Meshalkina, L.E.; Drutsa, V.L.; Koroleva, O.N.; Solovjeva, O.N.; Kochetov, G.A. Is transketolase-like protein, TKTL1, transketolase? Biochim. Biophys. Acta (Bba) Mol. Basis Dis. 2013, 1832, 387–390. [Google Scholar] [CrossRef] [Green Version]
  71. DeBerardinis, R.J.; Sayed, N.; Ditsworth, D.; Thompson, C.B. Brick by brick: Metabolism and tumor cell growth. Curr. Opin. Genet. Dev. 2008, 18, 54–61. [Google Scholar] [CrossRef] [Green Version]
  72. Fan, J.; Ye, J.; Kamphorst, J.J.; Shlomi, T.; Thompson, C.B.; Rabinowitz, J.D. Quantitative flux analysis reveals folate-dependent NADPH production. Nature 2014, 510, 298–302. [Google Scholar] [CrossRef] [Green Version]
  73. Snaebjornsson, M.T.; Janaki-Raman, S.; Schulze, A. Greasing the Wheels of the Cancer Machine: The Role of Lipid Metabolism in Cancer. Cell Metab. 2020, 31, 62–76. [Google Scholar] [CrossRef]
  74. Daniëls, V.W.; Smans, K.; Royaux, I.; Chypre, M.; Swinnen, J.V.; Zaidi, N. Cancer Cells Differentially Activate and Thrive on De Novo Lipid Synthesis Pathways in a Low-Lipid Environment. PLoS ONE 2014, 9, e106913. [Google Scholar] [CrossRef]
  75. Kuhajda, F.P. Fatty-acid synthase and human cancer: New perspectives on its role in tumor biology. Nutrition 2000, 16, 202–208. [Google Scholar] [CrossRef]
  76. Sabine, J.R.; Abraham, S.; Chaikoff, I.L. Control of Lipid Metabolism in Hepatomas: Insensitivity of Rate of Fatty Acid and Cholesterol Synthesis by Mouse Hepatoma BW7756 to Fasting and to Feedback Control. Cancer Res. 1967, 27, 793–799. [Google Scholar] [PubMed]
  77. Ookhtens, M.; Kannan, R.; Lyon, I.; Baker, N. Liver and adipose tissue contributions to newly formed fatty acids in an ascites tumor. Am. J. Physiol. Regul. Integr. Comp. Physiol. 1984, 247, R146–R153. [Google Scholar] [CrossRef] [PubMed]
  78. Swinnen, J.V.; Brusselmans, K.; Verhoeven, G. Increased lipogenesis in cancer cells: New players, novel targets. Curr. Opin. Clin. Nutr. Metab. Care 2006, 9, 358–365. [Google Scholar] [CrossRef] [PubMed]
  79. Ter Horst, K.W.; Serlie, M.J. Fructose Consumption, Lipogenesis, and Non-Alcoholic Fatty Liver Disease. Nutrients 2017, 9, 981. [Google Scholar] [CrossRef] [Green Version]
  80. Silva, J.C.; Marques, C.; Martins, F.O.; Viegas, I.; Tavares, L.; Macedo, M.P.; Jones, J.G. Determining contributions of exogenous glucose and fructose to de novo fatty acid and glycerol synthesis in liver and adipose tissue. Metab. Eng. 2019, 56, 69–76. [Google Scholar] [CrossRef]
  81. Zhao, S.; Jang, C.; Liu, J.; Uehara, K.; Gilbert, M.; Izzo, L.; Zeng, X.; Trefely, S.; Fernandez, S.; Carrer, A.; et al. Dietary fructose feeds hepatic lipogenesis via microbiota-derived acetate. Nature 2020, 579, 586–591. [Google Scholar] [CrossRef]
  82. Stanhope, K.L.; Bremer, A.A.; Medici, V.; Nakajima, K.; Ito, Y.; Nakano, T.; Chen, G.; Fong, T.H.; Lee, V.; Menorca, R.I.; et al. Consumption of Fructose and High Fructose Corn Syrup Increase Postprandial Triglycerides, LDL-Cholesterol, and Apolipoprotein-B in Young Men and Women. J. Clin. Endocrinol. Metab. 2011, 96, E1596–E1605. [Google Scholar] [CrossRef]
  83. Siemianowicz, K.; Gminski, J.; Stajszczyk, M.; Wojakowski, W.; Goss, M.; Machalski, M.; Telega, A.; Brulinski, K.; Magiera-Molendowska, H. Serum total cholesterol and triglycerides levels in patients with lung cancer. Int. J. Mol. Med. 2000. [Google Scholar] [CrossRef]
  84. Potischman, N.; McCulloch, C.E.; Byers, T.; Houghton, L.; Nemoto, T.; Graham, S.; Campbell, T.C. Associations between breast cancer, plasma triglycerides, and cholesterol. Nutr. Cancer 1991, 15, 205–215. [Google Scholar] [CrossRef]
  85. Ulmer, H.; Borena, W.; Rapp, K.; Klenk, J.; Strasak, A.; Diem, G.; Concin, H.; Nagel, G. Serum triglyceride concentrations and cancer risk in a large cohort study in Austria. Br. J. Cancer 2009, 101, 1202–1206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Sun, S.Z.; Empie, M.W. Fructose metabolism in humans - what isotopic tracer studies tell us. Nutr. Metab. 2012, 9, 89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Carreño, D.; Corro, N.; Torres-Estay, V.; Véliz, L.P.; Jaimovich, R.; Cisternas, P.; Francisco, I.F.S.; Sotomayor, P.C.; Tanasova, M.; Inestrosa, N.C.; et al. Fructose and prostate cancer: Toward an integrated view of cancer cell metabolism. Prostate Cancer Prostatic Dis. 2018, 22, 49–58. [Google Scholar] [CrossRef] [PubMed]
  88. O’Callaghan, B.L.; Koo, S.H.; Wu, Y.; Freake, H.C.; Towle, H.C. Glucose Regulation of the Acetyl-CoA Carboxylase Promoter PI in Rat Hepatocytes. J. Biol. Chem. 2001, 276, 16033–16039. [Google Scholar] [CrossRef] [Green Version]
  89. Rufo, C.; Teran-Garcia, M.; Nakamura, M.T.; Koo, S.H.; Towle, H.C.; Clarke, S.D. Involvement of a Unique Carbohydrate-responsive Factor in the Glucose Regulation of Rat Liver Fatty-acid Synthase Gene Transcription. J. Biol. Chem. 2001, 276, 21969–21975. [Google Scholar] [CrossRef] [Green Version]
  90. Wang, Y.; Botolin, D.; Xu, J.; Christian, B.; Mitchell, E.; Jayaprakasam, B.; Nair, M.; Peters, J.M.; Busik, J.; Olson, L.K.; et al. Regulation of hepatic fatty acid elongase and desaturase expression in diabetes and obesity. J. Lipid Res. 2006, 47, 2028–2041. [Google Scholar] [CrossRef] [Green Version]
  91. Crescenzo, R.; Bianco, F.; Falcone, I.; Coppola, P.; Liverini, G.; Iossa, S. Increased hepatic de novo lipogenesis and mitochondrial efficiency in a model of obesity induced by diets rich in fructose. Eur. J. Nutr. 2012, 52, 537–545. [Google Scholar] [CrossRef] [Green Version]
  92. Goncalves, M.D.; Lu, C.; Tutnauer, J.; Hartman, T.E.; Hwang, S.K.; Murphy, C.J.; Pauli, C.; Morris, R.; Taylor, S.; Bosch, K.; et al. High-fructose corn syrup enhances intestinal tumor growth in mice. Science 2019, 363, 1345–1349. [Google Scholar] [CrossRef]
  93. Liang, G.; Yang, J.; Horton, J.D.; Hammer, R.E.; Goldstein, J.L.; Brown, M.S. Diminished Hepatic Response to Fasting/Refeeding and Liver X Receptor Agonists in Mice with Selective Deficiency of Sterol Regulatory Element-binding Protein-1c. J. Biol. Chem. 2002, 277, 9520–9528. [Google Scholar] [CrossRef] [Green Version]
  94. Yamashita, H.; Takenoshita, M.; Sakurai, M.; Bruick, R.K.; Henzel, W.J.; Shillinglaw, W.; Arnot, D.; Uyeda, K. A glucose-responsive transcription factor that regulates carbohydrate metabolism in the liver. Proc. Natl. Acad. Sci. USA 2001, 98, 9116–9121. [Google Scholar] [CrossRef] [Green Version]
  95. Koo, H.Y.; Miyashita, M.; Cho, B.S.; Nakamura, M.T. Replacing dietary glucose with fructose increases ChREBP activity and SREBP-1 protein in rat liver nucleus. Biochem. Biophys. Res. Commun. 2009, 390, 285–289. [Google Scholar] [CrossRef] [PubMed]
  96. Softic, S.; Gupta, M.K.; Wang, G.X.; Fujisaka, S.; O’Neill, B.T.; Rao, T.N.; Willoughby, J.; Harbison, C.; Fitzgerald, K.; Ilkayeva, O.; et al. Divergent effects of glucose and fructose on hepatic lipogenesis and insulin signaling. J. Clin. Investig. 2017, 127, 4059–4074. [Google Scholar] [CrossRef] [Green Version]
  97. Li, M.V.; Chen, W.; Harmancey, R.N.; Nuotio-Antar, A.M.; Imamura, M.; Saha, P.; Taegtmeyer, H.; Chan, L. Glucose-6-phosphate mediates activation of the carbohydrate responsive binding protein (ChREBP). Biochem. Biophys. Res. Commun. 2010, 395, 395–400. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Ishimoto, T.; Lanaspa, M.A.; Le, M.T.; Garcia, G.E.; Diggle, C.P.; Maclean, P.S.; Jackman, M.R.; Asipu, A.; Roncal-Jimenez, C.A.; Kosugi, T.; et al. Opposing effects of fructokinase C and A isoforms on fructose-induced metabolic syndrome in mice. Proc. Natl. Acad. Sci. USA 2012, 109, 4320–4325. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Lanaspa, M.A.; Andres-Hernando, A.; Orlicky, D.J.; Cicerchi, C.; Jang, C.; Li, N.; Milagres, T.; Kuwabara, M.; Wempe, M.F.; Rabinowitz, J.D.; et al. Ketohexokinase C blockade ameliorates fructose-induced metabolic dysfunction in fructose-sensitive mice. J. Clin. Investig. 2018, 128, 2226–2238. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Todoric, J.; Caro, G.D.; Reibe, S.; Henstridge, D.C.; Green, C.R.; Vrbanac, A.; Ceteci, F.; Conche, C.; McNulty, R.; Shalapour, S.; et al. Fructose stimulated de novo lipogenesis is promoted by inflammation. Nat. Metab. 2020. [Google Scholar] [CrossRef]
  101. Stoiber, K.; Nagło, O.; Pernpeintner, C.; Zhang, S.; Koeberle, A.; Ulrich, M.; Werz, O.; Müller, R.; Zahler, S.; Lohmüller, T.; et al. Targeting de novo lipogenesis as a novel approach in anti-cancer therapy. Br. J. Cancer 2017, 118, 43–51. [Google Scholar] [CrossRef] [Green Version]
  102. Rysman, E.; Brusselmans, K.; Scheys, K.; Timmermans, L.; Derua, R.; Munck, S.; Veldhoven, P.P.V.; Waltregny, D.; Daniels, V.W.; Machiels, J.; et al. De novo Lipogenesis Protects Cancer Cells from Free Radicals and Chemotherapeutics by Promoting Membrane Lipid Saturation. Cancer Res. 2010, 70, 8117–8126. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Overview of fructose and glucose degradation pathways. Fructose uptake is mediated by GLUT5 and can be metabolized via different pathways: (1) phosphorylation via KHK to F1P, which is then hydrolyzed by aldolase B to DHAP and GA. Subsequently, fructose carbons can therefore enter glycolysis and the pentose phosphate pathway. (2) Phosphorylation via HK, entering glycolysis in the form of F6P, and (3) via the polyol pathway producing glucose which in turn can enter gylcolysis. Glucose can be taken up by several glucose transporters (here GLUT1) and is metabolized via glycolysis. Glucose can also be used for endogenous fructose production providing the possibility to bypass PFK. PFK is the rate-limiting enzyme of glycolysis and is inhibited by high levels of ATP and citrate as well as low pH and low oxygen levels. From an energetic point of view, the degradation pathways from fructose and glucose to pyruvate do not differ. Fructose degradation via KHK to GA3P consumes two ATP just as glucose in glycolysis. In the pay off phase, both pathways generate four ATP and two NADH+H+. Bypassing PFK inhibition by endogenous fructose production, however, consumes one additional NADPH+H+ and produces one NADH+H+. AKR1B1: aldo-keto reductase family 1, member B1; ALDO: aldolase; G3P: glycerol 3-phosphate; GA3P: glyceraldehyde 3-phosphate; GLUT1: glucose transporter 1; GLUT5: fructose transporter; GPDH: glycerol 3-phosphate dehydrogenase; GPI: glucose-6-phosphate isomerase; HK: hexokinase; KHK: keto hexokinase; PFK: phosphofructokinase; SORD: sorbitol dehydrogenase; TCA: tricarboxylic acid cycle; TK: triose kinase; TPI: triose-phosphate isomerase.
Figure 1. Overview of fructose and glucose degradation pathways. Fructose uptake is mediated by GLUT5 and can be metabolized via different pathways: (1) phosphorylation via KHK to F1P, which is then hydrolyzed by aldolase B to DHAP and GA. Subsequently, fructose carbons can therefore enter glycolysis and the pentose phosphate pathway. (2) Phosphorylation via HK, entering glycolysis in the form of F6P, and (3) via the polyol pathway producing glucose which in turn can enter gylcolysis. Glucose can be taken up by several glucose transporters (here GLUT1) and is metabolized via glycolysis. Glucose can also be used for endogenous fructose production providing the possibility to bypass PFK. PFK is the rate-limiting enzyme of glycolysis and is inhibited by high levels of ATP and citrate as well as low pH and low oxygen levels. From an energetic point of view, the degradation pathways from fructose and glucose to pyruvate do not differ. Fructose degradation via KHK to GA3P consumes two ATP just as glucose in glycolysis. In the pay off phase, both pathways generate four ATP and two NADH+H+. Bypassing PFK inhibition by endogenous fructose production, however, consumes one additional NADPH+H+ and produces one NADH+H+. AKR1B1: aldo-keto reductase family 1, member B1; ALDO: aldolase; G3P: glycerol 3-phosphate; GA3P: glyceraldehyde 3-phosphate; GLUT1: glucose transporter 1; GLUT5: fructose transporter; GPDH: glycerol 3-phosphate dehydrogenase; GPI: glucose-6-phosphate isomerase; HK: hexokinase; KHK: keto hexokinase; PFK: phosphofructokinase; SORD: sorbitol dehydrogenase; TCA: tricarboxylic acid cycle; TK: triose kinase; TPI: triose-phosphate isomerase.
Cells 09 02635 g001
Figure 2. Fructose contribution to the pentose phosphate pathway. Fructose can contribute to the oxidative (blue) as well as the non-oxidative (red) pentose phosphate pathway (PPP). To enter the oxidative PPP, fructose has to be either metabolized via the polyol pathway to glucose and further to G6P or directly via HK to F6P which can be converted to G6P. Direct phosphorylation via HK to F6P provides also a substrate for the non-oxidative PPP. Together with GA3P, F6P is converted by TKT to X5P and E4P. GA3P can also be provided by fructose metabolized via KHK. Both F6P and GA3P are used in further reactions of the non-oxidative PPP, demonstrating multiple entry points of fructose into the non-oxidative PPP. AKR1B1: aldo-keto reductase family 1, member B1; ALDO: aldolase; E4P: erythrose 4-phosphate; F6P: fructose 6-phosphate; G3P: glycerol 3-phosphate; GA3P: glyceraldehyde 3-phosphate; GLUT1: glucose transporter 1; GLUT5: fructose transporter; GPI: glucose-6-phosphate isomerase; HK: hexokinase; KHK: keto hexokinase; PFK: phosphofructokinase; SORD: sorbitol dehydrogenase; TCA: tricarboxylic acid cycle; TPI: triose-phosphate isomerase. TKT: transketolase; TALDO: transaldolase; G6PDH: glucose-6-phosphate deyhdrogenase; GSH: glutathione; GSSG: glutathione disulfide; 6PGL: 6-phosphogluconolactonase; 6PGDH: 6-phosphogluconate dehydrogenase; RPE: ribulose 5-phosphate 3-epimerase; RPI: ribose-5-phosphate isomerase; X5P: xylulose 5-phosphate.
Figure 2. Fructose contribution to the pentose phosphate pathway. Fructose can contribute to the oxidative (blue) as well as the non-oxidative (red) pentose phosphate pathway (PPP). To enter the oxidative PPP, fructose has to be either metabolized via the polyol pathway to glucose and further to G6P or directly via HK to F6P which can be converted to G6P. Direct phosphorylation via HK to F6P provides also a substrate for the non-oxidative PPP. Together with GA3P, F6P is converted by TKT to X5P and E4P. GA3P can also be provided by fructose metabolized via KHK. Both F6P and GA3P are used in further reactions of the non-oxidative PPP, demonstrating multiple entry points of fructose into the non-oxidative PPP. AKR1B1: aldo-keto reductase family 1, member B1; ALDO: aldolase; E4P: erythrose 4-phosphate; F6P: fructose 6-phosphate; G3P: glycerol 3-phosphate; GA3P: glyceraldehyde 3-phosphate; GLUT1: glucose transporter 1; GLUT5: fructose transporter; GPI: glucose-6-phosphate isomerase; HK: hexokinase; KHK: keto hexokinase; PFK: phosphofructokinase; SORD: sorbitol dehydrogenase; TCA: tricarboxylic acid cycle; TPI: triose-phosphate isomerase. TKT: transketolase; TALDO: transaldolase; G6PDH: glucose-6-phosphate deyhdrogenase; GSH: glutathione; GSSG: glutathione disulfide; 6PGL: 6-phosphogluconolactonase; 6PGDH: 6-phosphogluconate dehydrogenase; RPE: ribulose 5-phosphate 3-epimerase; RPI: ribose-5-phosphate isomerase; X5P: xylulose 5-phosphate.
Cells 09 02635 g002
Figure 3. De novo lipogenesis driven by fructose and glucose carbon. Both fructose and glucose can provide carbon for DNL. On the one hand, they can end up in citrate, which is then cleaved by ACLY to provide acetyl-CoA for fatty acid synthesis. On the other hand, they can provide G3P for triacylgyceride synthesis. Besides citrate, cytosolic acetyl-CoA can also be derived from acetate. It was shown that the microbiome of mice converts fructose to acetate. This microbiota-derived acetate is then taken up and utilized for DNL in mice [81]. ACC: acetyl-CoA carboxylase; ACSS2: acyl-coenzyme A synthetase short-chain family member 2; ACLY: ATP citrate lyase; AKG: α -ketoglutarate; FASN: fatty acid synthase; G3P: glycerol 3-phosphate; TCA: tricarboxylic acid cycle; TAG: triacylglyceride.
Figure 3. De novo lipogenesis driven by fructose and glucose carbon. Both fructose and glucose can provide carbon for DNL. On the one hand, they can end up in citrate, which is then cleaved by ACLY to provide acetyl-CoA for fatty acid synthesis. On the other hand, they can provide G3P for triacylgyceride synthesis. Besides citrate, cytosolic acetyl-CoA can also be derived from acetate. It was shown that the microbiome of mice converts fructose to acetate. This microbiota-derived acetate is then taken up and utilized for DNL in mice [81]. ACC: acetyl-CoA carboxylase; ACSS2: acyl-coenzyme A synthetase short-chain family member 2; ACLY: ATP citrate lyase; AKG: α -ketoglutarate; FASN: fatty acid synthase; G3P: glycerol 3-phosphate; TCA: tricarboxylic acid cycle; TAG: triacylglyceride.
Cells 09 02635 g003
Figure 4. Fructose signaling effects on DNL gene expression. Fructose intake activates ChREBP, which binds to ChRE, a promotor of lipogenic emzymes. Furthermore, fructose is also involved in SREBF-mediated DNL gene expression. SREBF is activated by insulin signaling, binds to SRE, promoting gene expression of lipogenic enzmyes. Increased fructose consumption results in insulin resistance, which in turn increases insulin secretion in a compensatory manner. Therefore, fructose consumption affects DNL gene expression by ChREBP and SREBF signaling. ACC: acetyl-CoA carboxylase; ACSS2: acyl-coenzyme A synthetase short-chain family member 2; ACLY: ATP citrate lyase; AKG: α -ketoglutarate; ChREBP: carbohydrate-responsive element-binding protein; ChoRE: carbohydrate response element; DNL: de novo lipogenesis; FASN: fatty acid synthase; G3P: glycerol 3-phosphate; SRE: sterol regulatory element; SREBF: sterol regulatory element-binding protein; TCA: tricarboxylic acid cycle; TAG: triacylglyceride.
Figure 4. Fructose signaling effects on DNL gene expression. Fructose intake activates ChREBP, which binds to ChRE, a promotor of lipogenic emzymes. Furthermore, fructose is also involved in SREBF-mediated DNL gene expression. SREBF is activated by insulin signaling, binds to SRE, promoting gene expression of lipogenic enzmyes. Increased fructose consumption results in insulin resistance, which in turn increases insulin secretion in a compensatory manner. Therefore, fructose consumption affects DNL gene expression by ChREBP and SREBF signaling. ACC: acetyl-CoA carboxylase; ACSS2: acyl-coenzyme A synthetase short-chain family member 2; ACLY: ATP citrate lyase; AKG: α -ketoglutarate; ChREBP: carbohydrate-responsive element-binding protein; ChoRE: carbohydrate response element; DNL: de novo lipogenesis; FASN: fatty acid synthase; G3P: glycerol 3-phosphate; SRE: sterol regulatory element; SREBF: sterol regulatory element-binding protein; TCA: tricarboxylic acid cycle; TAG: triacylglyceride.
Cells 09 02635 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Krause, N.; Wegner, A. Fructose Metabolism in Cancer. Cells 2020, 9, 2635. https://doi.org/10.3390/cells9122635

AMA Style

Krause N, Wegner A. Fructose Metabolism in Cancer. Cells. 2020; 9(12):2635. https://doi.org/10.3390/cells9122635

Chicago/Turabian Style

Krause, Nils, and Andre Wegner. 2020. "Fructose Metabolism in Cancer" Cells 9, no. 12: 2635. https://doi.org/10.3390/cells9122635

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

Article Metrics

Back to TopTop