Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

ll

Review
Molecular aspects of fructose
metabolism and metabolic disease
Mark A. Herman1,2,3,* and Morris J. Birnbaum4,*
1Division of Endocrinology, Metabolism, and Nutrition, Duke University, Durham, NC, USA
2Duke Molecular Physiology Institute, Duke University, Durham, NC, USA
3Department of Pharmacology and Cancer Biology, Duke University, Durham, NC, USA
4Internal Medicine Research Unit, Pfizer, Cambridge, MA, USA

*Correspondence: mark.herman@duke.edu (M.A.H.), morris.birnbaum@pfizer.com (M.J.B.)


https://doi.org/10.1016/j.cmet.2021.09.010

SUMMARY

Excessive sugar consumption is increasingly considered as a contributor to the emerging epidemics of


obesity and the associated cardiometabolic disease. Sugar is added to the diet in the form of sucrose or
high-fructose corn syrup, both of which comprise nearly equal amounts of glucose and fructose. The unique
aspects of fructose metabolism and properties of fructose-derived metabolites allow for fructose to serve as
a physiological signal of normal dietary sugar consumption. However, when fructose is consumed in excess,
these unique properties may contribute to the pathogenesis of cardiometabolic disease. Here, we review the
biochemistry, genetics, and physiology of fructose metabolism and consider mechanisms by which exces-
sive fructose consumption may contribute to metabolic disease. Lastly, we consider new therapeutic options
for the treatment of metabolic disease based upon this knowledge.

INTRODUCTION has correlated with an explosion in the prevalence of cardiome-


tabolic diseases, refocusing public health attention on this issue
The fixation of carbon dioxide into sugars via photosynthesis (Bray et al., 2004; Malik et al., 2010).
forms the basis of Earth’s food web. Heterotrophs cannot syn- Sugar is added to beverages and food products in the form of
thesize their own sugars; therefore, they evolved to use plant- sucrose or the industrial product high-fructose corn syrup
derived sugars for energetic and synthetic purposes. Sucrose, (HFCS), which is manufactured by digestion of corn starch to
the predominant circulating sugar in plants, is composed of its glucose monosaccharides by microbial enzymes. Microbial
equal parts of glucose and fructose. Glucose is the major circu- xylose isomerase then catalyzes the conversion of some portion
lating sugar in animals, and the amount of fructose is negligible in of the glucose monosaccharides to fructose. ‘‘HFCS 55,’’ the
comparison. Glucose is a primary energetic and synthetic fuel for most common formulation added to beverages, consists of
most tissues and cell types in the body. In contrast, the fructose 55% fructose and 45% glucose. Sucrose is ingested as a disac-
component of the ingested sucrose is rapidly cleared by the in- charide and requires cleavage to its monosaccharide constitu-
testines and liver and is catabolized for energetic purposes, con- ents, glucose and fructose, prior to absorption. This is catalyzed
verted to glucose and its polymeric storage form, glycogen, or to by the enzyme sucrase-isomaltase, which is localized to the in-
fatty acids and is stored as triglycerides. Despite its low levels in testinal lumen’s brush-border membrane (Hauri et al., 1979). In
the circulation, fructose serves as a signal of the ingested dietary contrast, the sugars in HFCS are ingested as monosaccharides
sugar. The differences by which glucose and fructose are sensed and require no further processing prior to absorption. The differ-
and metabolized are of fundamental importance to the re- ences in monosaccharide versus disaccharide content of HFCS
sponses to normal physiological sugar consumption as well as versus sucrose could differentially affect the rate of absorption of
the pathophysiological consequences of excessive ingestion. these sugars, which might result in distinct biological effects (Le
We are amid a major emergence of obesity, diabetes, and et al., 2012). However, compelling data that sucrose and HFCS
associated cardiometabolic diseases, including non-alcoholic impart distinct metabolic effects on humans are lacking (Stan-
fatty liver disease (NAFLD). The etiology of these epidemics is hope et al., 2008).
multifactorial and depends on the interactions between genetics The contribution of natural dietary sugar and added sugars to
and environmental factors, including diet and physical activity. the epidemics of cardiometabolic disease remains controversial
Concerns that excessive consumption of sugary foods and bev- (Khan and Sievenpiper, 2016; Stanhope, 2016; Ter Horst and
erages might contribute to the development of obesity and dia- Serlie, 2017). The epidemiological correlations between different
betes date back as far as the first millennium BCE and remerge sugar exposures and outcomes, including energy intake, weight
periodically throughout modern history (Emerson and Larimore, gain, type 2 diabetes (T2D), dyslipidemia, insulin resistance, hy-
1924; Johnson et al., 2017). In the last half century, enhanced in- pertension, hyperuricemia, and dental caries, tend to be less
dustrial processes and corn subsidies increased the availability consistent for measurements of total sugars and more consis-
of sugar, particularly in the form of sweetened beverages. This tent for measurements of free and added sugars (Dhingra

Cell Metabolism 33, December 7, 2021 ª 2021 Elsevier Inc. 2329


ll
Review

et al., 2007; Green et al., 2014; Haslam et al., 2020; Imamura they learn to prefer foods that are high in glucose content, but not
et al., 2015; Jayalath et al., 2015; Khan and Sievenpiper, 2016; fructose or other noncaloric sweeteners (de Araujo et al., 2008;
de Koning et al., 2012; Mela and Woolner, 2018). Consumption Sclafani et al., 2014). This form of sugar preference requires
of fruit, the major natural source of dietary sugar, is associated the intestinal sodium-glucose cotransporter (SLC5A1, also
with healthy outcomes, although fruit consumption is often asso- known as SGLT1), the principal transporter expressed by both
ciated with other healthy lifestyle choices (Khan and Sievenpiper, enterocytes and enteroendocrine cells that is responsible for
2016). Sugar-sweetened beverages (SSBs) are the largest con- the uptake of glucose and galactose across the luminal intestinal
tributors to added sugar and fructose intake in children and membrane (Dyer et al., 2005; Gorboulev et al., 2012; Tan et al.,
adults (U.S. Department of Agriculture and U.S. Department of 2020). The SGLT1-dependent effects on sugar consumption
and Health and Human Services, 2020; Virani et al. 2020), and are mediated by signals carried by the vagus nerve to activate
their consumption consistently associates with increased cardi- neurons in the caudal nucleus of the solitary tact, a region previ-
ometabolic risk (Dhingra et al., 2007; Green et al., 2014; Haslam ously implicated in transducing gut sensations to the brain (Han
et al., 2020; de Koning et al., 2012; Ma et al., 2015, 2016; Malik et al., 2018; Kaelberer et al., 2018; Tan et al., 2020). However,
et al., 2013; McKeown et al., 2018; Xi et al., 2015; Yoshida recent work in an alternative ‘‘tasteless’’ genetic mouse model
et al., 2007; Zhang et al., 2020), including the excess cardiome- missing the P2X2 and P2X3 purinergic receptors that are
tabolic deaths associated with poor diet (Coffee and Caffeine required for taste cell signaling to gustatory nerves suggests
Genetics Consortium et al., 2015). alternative forms of post-ingestive sugar sensing (Andres-Her-
Recommendations from public health organizations and pol- nando et al., 2020a; Sclafani and Ackroff, 2021). These studies
icy makers to reduce sugar consumption have met with modest show that fructose can enhance sugar consumption indepen-
success over the last decade. Approximately 50% of adults and dent of taste. The combination of sweet taste sensation and
100% of youths report consuming at least one SSB per day, a post-ingestive sugar sensing likely interacts to robustly reinforce
level of sugar consumption that is associated with adverse car- preferences for sugar-rich foods.
diometabolic risk factors (Bailey et al., 2018; Rosinger et al.,
2017; Virani et al., 2020). Moreover, consumption of added EXCESSIVE FRUCTOSE CONSUMPTION CAN CAUSE
sugars, even at the upper limits of recent dietary recommenda- METABOLIC DISEASE
tions, may adversely impact health (Stanhope et al., 2011). Pop-
ulations with greater food insecurity and certain ethnic or racial In the 1960s, observations that diets high in fructose can induce
subgroups report even higher intakes of SSB (Virani et al., hypertriglyceridemia in animals and humans in a matter of days
2020; Zagorsky and Smith, 2020). SSB consumption has much more robustly than diets containing comparable amounts
increased during the COVID-19 pandemic (Flanagan et al., of starch or glucose led to interest in fructose as a contributor to
2021). Although public health efforts to reduce sugar consump- metabolic disease (Macdonald, 1966; Nikkila € and Ojala, 1965).
tion are warranted, an improved understanding of the mecha- Carbohydrate- and sugar-induced hypertriglyceridemia was
nisms by which SSBs and their constituent sugars contribute also strongly associated with hyperinsulinemia in humans and
to disease is paving the way for new therapeutic strategies. in animal models of obesity and overnutrition (Eaton and Kipnis,
These mechanisms and strategies will be the focus of this 1969; Farquhar et al., 1966; Reaven et al., 1979; Reiser and Hall-
review. frisch, 1977). Subsequent studies showed that very-high-fruc-
tose diets could induce numerous dysmetabolic features,
SWEET TASTE PERCEPTION including hypertension, that were typical of the ‘‘metabolic syn-
drome,’’ more recently known as a syndrome of cardiometabolic
Animals have evolved complex mechanisms to sense sugar and disease (Hwang et al., 1987; Reaven, 1988; Sleder et al., 1980;
motivate its consumption. Monosaccharides, including glucose Tobey et al., 1981). Although these studies were instrumental
and fructose, as well as disaccharides, such as sucrose, potently in demonstrating the potential for fructose as a cause of meta-
activate the G protein-coupled receptors Tas1R2 and Tas1R3, bolic disease, they often used fructose exposures that exceeded
which are located on the epithelial cells of the tongue and palate 50% of the total energy, far exceeding typical consumption by
(Nelson et al., 2001; Zhao et al., 2003). These receptors stimulate humans. Fructose accounts for 9% of energy intake in the
neuronal circuitry, projecting to regions of the amygdala involved US, and even consumers in the 95th percentile consume
in processing hedonic and aversive stimuli (Wang et al., 2018). 15% of their energy as fructose (Marriott et al., 2009). Never-
Some investigators have proposed that sugar has addictive theless, diets unphysiologically high in fructose may be useful
properties by modulating mesolimbic dopaminergic reward cir- for identifying molecular mechanisms relevant for common
cuitry (Avena et al., 2008; Leigh and Morris, 2018; Lustig, forms of diet-induced metabolic diseases.
2010). Indeed, the sweetness of both caloric and noncaloric In recent decades, interventional studies spanning weeks to
sweeteners can induce or increase the consumption of neutral months have shown that overfeeding moderate to high doses
or aversive substances, such as alcohol, in strains of mice that of sugars containing fructose can adversely impact metabolic
otherwise avoid it (Yoneyama et al., 2008). Nevertheless, outcomes in humans. In the longest of such an interventional
compelling data for the ‘‘addictive’’ potential of sugar in humans study, Maersk and colleagues found that adding one liter of
are lacking (Leigh and Morris, 2018; Markus et al., 2017). SSB daily for 6 months increased visceral, liver, and ectopic
Mechanisms in addition to sweet taste contribute to the moti- fat (Maersk et al., 2012). These increases were not observed in
vation for sugary food consumption. Mice with deficiency of both those consuming isocaloric skim milk, noncaloric diet, soda, or
Tas1R2 and Tas1R3 cannot taste sugar or other sweeteners, yet water (Maersk et al., 2012). Adding SSBs ranging from 10% to

2330 Cell Metabolism 33, December 7, 2021


ll
Review

25% of energy requirements dose-dependently increased car- 2011). Fructose supplementation in non-human primates can
diovascular risk factors, including lipids and uric acid, over a induce dyslipidemia, steatosis, and the progression of NAFLD
2-week period (Stanhope et al., 2015). (Butler et al., 2019, 2020; Cydylo et al., 2017; Kavanagh et al.,
With respect to potential adverse effects of fructose, Taskinen 2013) and may provide a more faithful model of the pathogenesis
et al. reported that a moderate increase in fructose consumption of human cardiometabolic disease.
(75 g or 300 kcal per day) in men with obesity for 12 weeks Although the literature largely focuses on the health risks of
increased body weight, liver fat, hepatic de novo lipogenesis increasing amounts of sugar, the mode of sugar exposure may
(DNL), and other cardiovascular risk factors (Taskinen et al., also affect its biological impact. In animal studies, sugar pro-
2017). A pivotal study by Stanhope et al. compared the effects vided in liquid form may be more deleterious than when incorpo-
of adding isocaloric glucose- and fructose-sweetened bever- rated into solid food (Jang et al., 2020; Togo et al., 2019).
ages, constituting 25% of the basal energy requirement over Additionally, sugar ingested as a single large daily bolus may
10 weeks in adults who were overweight or obese (Stanhope be more detrimental than frequent ingestion of smaller amounts
et al., 2009). Both interventions increased body weight to a of fructose (Jang et al., 2020). These variables are not typically
similar degree. However, fructose, but not glucose, increased assessed in current epidemiological studies and may contribute
visceral adiposity, DNL, an atherogenic dyslipidemia, and to the variable associations of dietary sugar exposures with car-
indices of insulin resistance (Stanhope et al., 2009). Whereas diometabolic health.
this study provides evidence that added fructose is more delete- Reduced physical activity is another environmental change
rious than added glucose, a recent report from the same group that is often invoked to explain our current obesity and diabetes
revealed that fructose and glucose in combination may be epidemics. Exercise acutely increases energy requirements and
more harmful than fructose alone (Hieronimus et al., 2020). Fruc- mobilizes metabolic substrates to support muscle contraction.
tose supplementation to 25% of the energy requirement pro- Based on evidence that fructose provides some energetic
duced the strongest effect on increasing circulating triglycerides, advantage when supplied during exercise, Tappy and Rosset
whereas the combination of fructose and glucose in HFCS eli- speculated that the adverse effects of fructose are magnified
cited a greater increase in LDL-cholesterol and circulating apoli- in sedentary people (Tappy, 2018; Tappy and Rosset, 2017).
poprotein B levels. This result demonstrates the potential for Spontaneous running attenuates sucrose-induced hypertrigly-
complex interactions between distinct sugars. Similar interac- ceridemia in rats (Zavaroni et al., 1981). Modern hunter-gath-
tions also likely occur between sugars and other dietary macro- erers of the Hadza tribe consume from 8% to 16% of their energy
nutrients, including fat and protein, to impact metabolic health intake as honey, which is largely fructose and glucose (Pontzer
(Softic et al., 2018). et al., 2012). This is similar to the 50th and 95th percentiles of
The most compelling clinical evidence that dietary sugar, sugar consumers, respectively, in the Western societies. Yet
when consumed in amounts typical of Western diets, contributes Hadza foragers have a low prevalence of cardiovascular risk fac-
to adverse metabolic health comes from restriction studies, tors (Pontzer et al., 2012; Raichlen et al., 2017). Their high levels
particularly those performed in children and adolescents. of physical activity may mitigate the adverse effects of sugar
Some, but not all, studies showed that interventions aimed at consumption.
reducing SSB consumption reduce weight gain, adiposity, liver Because of the growing concerns regarding the role of sugar
fat, and indices of insulin resistance, depending on the study consumption in cardiometabolic disease, public health organiza-
population, the degree of sugar restriction and the duration of tions have invested in measures aimed at reducing sugar con-
the study (Ebbeling et al., 2012; de Ruyter et al., 2012; Schwarz sumption, including dietary recommendations concerning
et al., 2017; Schwimmer et al., 2019). ‘‘safe’’ limits for sugar consumption (Johnson et al., 2009; U.S.
Although the evidence that added sugars containing fructose Department of Agriculture and U.S. Department of and Health
can negatively impact health indices is compelling, whether and Human Services, 2020; World Health Organization 2015).
these adverse effects are mediated exclusively through the ef- Given the lack of consensus regarding health risks and safe
fects of increased energy intake to enhance weight gain or thresholds, recommendations regarding sugar consumption
also through mechanisms independent of adiposity is less clear. vary substantially. Aside from dietary recommendations, public
Complexity, cost, and ethical issues limit large-scale, long-term health studies have largely focused on behavioral interventions
dietary interventions that are needed to confirm effects on meta- to reduce SSB and sugar consumption in at-risk school-age chil-
bolic and cardiovascular outcomes and to fully disentangle the dren with modest benefits (Rahman et al., 2018; Vercammen
effects of added sugar on weight gain and adiposity from other et al., 2018). Interventions aimed at altering the physical or social
deleterious effects. Because of the difficulty in conducting environment to reduce the selection of SSBs, such as warning la-
long-term dietary studies in ‘‘free-range’’ humans, studies in an- bels, in-store promotions of healthier beverages, price increases
imal models remain valuable. Non-human primate models with on SSBs, government food benefits programs to disincentivize
physiologies that closely approximate our own may be an impor- SSB purchases, and increasing the availability of low-calorie
tant alternative model. Indeed, supplementation of the diets of beverages in the home and community campaigns targeting
rhesus macaques with 300 kcal per day of fructose-sweetened SSBs, have all shown some success (von Philipsborn et al.,
beverages, similar to that used in the Taskinen et al. study, for 2020). Population-scale efforts to reduce SSB consumption
up to 1 year produced many features of the metabolic syndrome, through SSB taxes appear effective (Fernandez and Raine,
including increased body weight, fat mass, insulin resistance, 2019; Redondo et al., 2018). However, whether any of these ef-
dyslipidemia with hypertriglyceridemia and decreased HDL forts enhance health outcomes remains uncertain (Pfinder
cholesterol, and in some cases, overt diabetes (Bremer et al., et al., 2020).

Cell Metabolism 33, December 7, 2021 2331


ll
Review

INTESTINAL FRUCTOSE ABSORPTION major role in transcellular monosaccharide efflux (Stu €mpel
et al., 2001).
Following ingestion, fructose travels the gastrointestinal tract At birth, the expression of intestinal GLUT5 is negligible and in-
to the small intestine where it is absorbed via transporters ex- creases at weaning (Douard and Ferraris, 2008; Ferraris, 2001).
pressed on the brush border of intestinal epithelial cells. This induction is enhanced when such diets include fructose
GLUT5 (also known as Slc2a5) has marked specificity for fruc- (Ferraris, 2001). Fructose-mediated induction of intestinal
tose over glucose, although a wide range of Km for fructose (6 GLUT5 expression is dependent on intestinal fructose meta-
to 15 mM) has been reported depending on the assay system bolism (Patel et al., 2015b). The induction of intestinal GLUT5
(Burant et al., 1992; Douard and Ferraris, 2008). It is expressed expression in response to dietary fructose is mediated by carbo-
at high levels on the luminal membrane of the small intestine hydrate-responsive element-binding protein (ChREBP, also
with lesser expression on the basolateral membrane (Burant known as Mlxipl), a carbohydrate-sensing transcription factor
et al., 1992; Davidson et al., 1992; Ferraris et al., 2018; Rand expressed at high levels in the intestinal epithelia and other key
et al., 1993). Its expression and activity are highest in the prox- metabolic tissues (Kato et al., 2018; Kim et al., 2017; Oh et al.,
imal duodenum and decline along the length of the small intes- 2018). A critical role for ChREBP in organismal fructose meta-
tine (Rand et al., 1993). GLUT5 is essential for the intestinal bolism was demonstrated by observations that in mice with
absorption of dietary fructose. GLUT5 knockout mice are global or intestine-specific ChREBP knockout, exposure to
healthy and fertile on standard rodent chow diets, which high-fructose diets failed to induce intestinal GLUT5 expression,
contain little or no fructose (Barone et al., 2009). However, resulting in a malabsorption syndrome and morbidity within
when exposed to diets containing fructose, these mice 1 week (Iizuka et al., 2004; Kato et al., 2018; Kim et al., 2017;
develop a severe, morbid malabsorption syndrome character- Oh et al., 2018). This morbidity is similar to the malabsorption
ized by diarrhea and intestinal distension (Barone et al., 2009) phenotype observed in GLUT5 knockout mice (Barone
(Table 1). et al., 2009).
GLUT5 is a facilitative transporter; therefore, fructose trans- Fructose absorption is highly variable in human children and
port into the enterocyte is proportional to the concentration adults, depending on age, dietary sugar exposure, and likely
gradient across the enterocyte luminal membrane. This con- other unidentified genetic, dietary, and environmental factors
trasts with the intestinal glucose absorption, which is mediated (reviewed in Ferraris et al., 2018). Fruit juices, such as apple juice,
by the sodium-linked cotransporter, SGLT1, which can effi- which contain large amounts of fructose, are commonly
ciently pump glucose into the enterocyte against a concentration consumed in large quantities by young children and may cause
gradient. This explains, in part, why intestinal glucose absorption nonspecific gastrointestinal symptoms that resolve with reduc-
is faster and more complete than that of fructose (Cori, 1925). tions in juice consumption (Ushijima et al., 1995). Fructose
Robust fructose phosphorylation within the enterocyte may be malabsorption can be measured via detection of colonic flora-
important to maintain a steep luminal-enterocyte fructose derived gases, such as hydrogen or methane, after ingestion of
gradient to facilitate more complete fructose absorption (Patel an oral fructose load (Rao et al., 2007). Some, but not all, studies
et al., 2015a). In contrast with glucose, which is nearly fully using such methods find associations between fructose malab-
absorbed within the small intestine, the intestinal capacity for sorption and functional gastrointestinal symptoms (Fernández-
fructose transport can be saturated and some proportion of an Bañares et al., 1993; Melchior et al., 2014). Reducing the
ingested fructose load may escape the small intestine and reach consumption of dietary fructose, in addition to other highly
the large intestine, where it is readily catabolized by gut micro- fermentable but poorly absorbed carbohydrates and polyols,
biota (Zhao et al., 2020). may be useful to treat a range of gastrointestinal symptoms
Although GLUT5 expression is responsible for intestinal fruc- and conditions (Gibson and Shepherd, 2005).
tose absorption, it is also expressed in other tissues and cell Although genetic ablation of GLUT5 causes fructose malab-
types, including skeletal muscle, pre-adipocytes, the prostate sorption in mice, fructose malabsorption in humans is not neces-
gland, spermatozoa, and erythrocytes (Burant et al., 1992; sarily associated with reduced intestinal GLUT5 expression
Concha et al., 1997; Kayano et al., 1990; Reinicke et al., 2012). (Wasserman et al., 1996). Thioredoxin-interacting protein
However, the importance of fructose transport into these tissues (TXNIP) is another ChREBP-regulated, fructose-induced protein
remains unclear. that is expressed in key metabolic tissues, including the liver and
Mechanisms mediating the efflux of both glucose and fructose intestine (Dotimas et al., 2016; Iizuka et al., 2004). TXNIP facili-
across the enterocyte basolateral membrane are less certain. tates the localization of facilitative hexose transporters to plasma
Both glucose and fructose efflux may be facilitated by GLUT2 membranes, including localization of GLUT5 to enterocyte
(also known as Slc2a2), which is expressed at high levels on luminal membranes (Shah et al., 2020). Thus, factors regulating
basolateral membranes in enterocytes of the proximal small in- proper GLUT5 localization and function, in addition to its expres-
testine (Davidson et al., 1992; Ferraris et al., 2018). GLUT2 is a sion, may also be important determinants of dietary fructose ab-
low-affinity, high-capacity facilitative transporter for both sorption.
glucose and fructose with similar affinities for both sugars (Km It is not known whether natural variation in intestinal fructose
11 mM) (Manolescu et al., 2007). However, glucose efflux absorption or hepatic fructose metabolism contributes to fruc-
does not appear to be markedly impaired in either GLUT2-defi- tose-induced cardiometabolic disease. Walker and colleagues
cient humans or mice (Santer et al., 2003; Stu €mpel et al., noted that African Americans had higher rates of fructose
2001). As an alternative, exocytosis via a microsomal mem- malabsorption associated with lower liver fat compared with His-
brane-based transport pathway has been proposed to play a panic controls and that within African Americans, fructose

2332 Cell Metabolism 33, December 7, 2021


Review
Table 1. Metabolic effects resulting from genetic manipulation of fructolytic enzymes and related factors in animals challenged with high dietary fructose
Gene target Intervention Phenotype References
GLUT5 (SLC2A5) global KO malabsorption syndrome associated with Barone et al., 2009
intestinal distension, diarrhea, and marked
weight loss within 1 week
KHKa/c global KO reduced weight gain and adiposity; reduced Diggle et al., 2009;
glycemia and insulin; reduced hepatic Ishimoto et al., 2012
steatosis; mildly reduced ad libitum
fructose consumption; increased urinary
fructose excretion
liver KD – ASO reduced weight gain; reduced hepatic Softic et al., 2018
steatosis; improved glucose tolerance
liver KO reduced weight gain; reduced hepatic Andres-Hernando et al., 2020b
steatosis; reduced insulin; increased
urinary fructose excretion
intestine KO reduced ad libitum fructose consumption; Andres-Hernando et al., 2020b
despite reduced fructose consumption,
persistent increases in body weight and
hepatic steatosis
KHKa global KO increased weight gain and adiposity; (Diggle et al., 2010);
increased insulin and hepatic steatosis Ishimoto et al., 2012
KHKc intestine KO increased portal fructose levels; increased Jang et al., 2020
fructose delivery to microbiota; increased
hepatic DNL; increased circulating
triglycerides and hepatic steatosis
intestine OX reduced portal fructose levels; reduced Jang et al., 2020
Cell Metabolism 33, December 7, 2021 2333

DNL, reduced ad libitum fructose


consumption
ALDOB global KO steatosis on fructose-free diets and rapid Lanaspa et al., 2018a;
development of severe steatosis and Oppelt et al., 2015
inflammation with fructose exposure;
morbidity and mortality within 1 week of
fructose exposure; rescue with
pharmacological inhibition of KHK
TKFC global KO – Liu et al., 2020
(Continued on next page)

ll
2334 Cell Metabolism 33, December 7, 2021

ll
Table 1. Continued
Gene target Intervention Phenotype References
ChREBP (MLXIPL) global KO weight loss, hypothermia, and severe Iizuka et al., 2004
morbidity and mortality within 1 week of
fructose challenge; increased hepatic
glycogen and hexose phosphates
global KO prevents fructose-induced expression of Kim et al., 2016
the hepatic fructolytic program; reduced
endogenous glucose production; impaired
glycogenolysis; increased hepatic hexose
phosphates.
global KO fructose and sucrose malabsorption Kato et al., 2018;
associated with diarrhea, intestinal Oh et al., 2018
distension, and alterations in microbiome
global KO weight loss; liver inflammation (Zhang et al., 2017)
liver KD – ASO (rat) reduced circulating triglycerides; enhanced Erion et al., 2013
peripheral insulin sensitivity; reduced DNL;
increased hepatic glycogen; mild
transaminitis without liver inflammation
liver KO reduced weight gain and adiposity; reduced Kim et al., 2017
circulating insulin; reduced DNL; increased
hepatic glycogen; preserved glycerol
tolerance; mild transaminitis without liver
inflammation
liver KO marked hepatic glycogen accumulation (Shi et al., 2020)
associated with transaminitis without liver
inflammation
intestine KO weight loss with fructose malabsorption Kim et al., 2017
and intestinal distension associated with
severe morbidity and mortality within
1 week of fructose challenge
KO, knockout; KD, knockdown; OX, overexpression. All genetic interventions were performed in mice unless otherwise specified.

Review
ll
Review
Figure 1. Organismal fructose metabolism
and first-pass extraction in the gut
Following ingestion of a large oral fructose load,
fructose is absorbed into intestinal epithelial enter-
ocytes. A portion of this fructose is phosphorylated
by KHK within the enterocyte and is converted to
glucose, lactate, glycerate, and other organic acids,
which travel via the portal vein to the liver. Portal
fructose concentrations can transiently reach con-
centrations as high as 1 mM. Fructose reaching
the liver is efficiently extracted by hepatocytes and
phosphorylated by KHK, where it can be used for
glucose production, lipogenesis, glycogen synthe-
sis, and energetic purposes. Peripheral blood
fructose concentrations transiently peak at levels
10-fold lower than peak portal levels.

This perspective originated from classical


studies conducted by Denker et al., Holds-
worth et al., and others that measured 1
malabsorption correlated inversely with liver fat (Walker et al., to 2 mM fructose in the portal vein of human study participants
2012). Another small study indicated that children with NAFLD following oral administration of fructose (Cook, 1969; Dencker
may have enhanced fructose absorption compared with lean con- et al., 1972; Holdsworth and Dawson, 1965). Cotemporaneous
trols and that children with either NAFLD or obesity may have work in baboons demonstrated that the gastric installation of 2
enhanced fructose metabolism compared with their lean counter- g/kg body mass sucrose achieved fructose concentrations of
parts (Sullivan et al., 2015). Intriguingly, recent work has shown 2 mM in the portal vein, whereas simultaneous measurements
that GLUT5 expression and glucose production are increased in femoral arterial blood transiently peaked at 0.7 mM (Crossley
from intestinal stem-cell-derived enteroids generated from intes- and Macdonald, 1970). Similar results were obtained in other an-
tinal biopsies obtained from patients who had obesity compared imals (Topping and Mayes, 1971). The large decrement in fruc-
with lean controls (Hasan et al., 2021). Although these studies tose concentrations from portal blood to peripheral arterial blood
suggest the possibility that innate differences in intestinal function speaks to the efficiency of hepatic fructose extraction confirmed
might contribute to fructose-related metabolic phenotypes, it re- in the liver perfusion experiments (Mayes, 1993).
mains unclear whether these observed differences might be Recent work in mice conducted by Jang et al. refocused atten-
causes or consequences of obesity and NAFLD. tion on the relative importance of liver versus intestinal fructose
metabolism (Jang et al., 2018). Using a stable isotope approach,
ORGANISMAL FRUCTOSE METABOLISM AND THE Jang and colleagues reported that in mice gavaged with fructose
IMPORTANCE OF THE INTESTINE AND LIVER at a dose of 0.5 g/kg body mass, 90% of the absorbed fructose
was metabolized within the small intestine and a relatively little
The innate differences in mammalian glucose versus fructose amount reached the liver. Most of the fructose metabolized
metabolism are reflected in the marked differences in their circu- within the intestine appeared in the portal vein as glucose or
lating blood levels. Whereas normal fasting glucose concentra- other fructose-derived metabolic products, including lactate,
tions in peripheral blood are 5 mM, fructose circulates at less alanine, glycerate, and other organic acids. The small intestine
than 0.02 mM in fasted conditions (Chen et al., 2020; Francey expresses the full complement of gluconeogenic enzymes and
et al., 2019). The ingestion of a large, isolated glucose load as is capable of robust conversion of ingested fructose into circu-
performed during a glucose tolerance test produces peripheral lating glucose (Bismut et al., 1993; Ginsburg and Hers, 1960;
glycemia of 7.5 mM in healthy individuals. In contrast, an equiv- Jang et al., 2018). Although studies show that the robust conver-
alently large oral fructose load transiently increases the periph- sion of fructose to glucose is readily detectable in the intestines
eral circulating fructose levels with peaks rarely exceeding of many species, earlier studies suggested that this did not occur
1 mM. This is followed by a rapid return to low micromolar in humans (Cook, 1969; Holdsworth and Dawson, 1965). Howev-
levels within hours. The absence of a comparatively large in- er, reexamination of isotopic tracer data in studies performed in
crease in the peripheral circulating fructose following ingestion human children is consistent with extensive conversion of in-
of a large oral fructose load illustrates the gut’s ability to metab- gested fructose to glucose in human intestines as well (Bismut
olize most of the orally ingested fructose. Relatively little escapes et al., 1993; Gopher et al., 1990).
to the peripheral circulation (Figure 1). Using a sophisticated Whether intestinal fructose metabolism is dominant over liver
dual-tracer approach in humans, Fancey et al. recently demon- fructose metabolism in animals other than mice is less clear.
strated that after ingesting a drink containing 30 g fructose and The relative importance of intestinal fructose metabolism may
30 g glucose, only 15% of the fructose escaped the first-pass be different among species (Ginsburg and Hers, 1960). The in-
gut metabolism (Francey et al., 2019). testinal GLUT5 expression is markedly higher in rats and humans
For the last half century, conventional wisdom suggested that compared with mice (Kim et al., 2007). These differences may
the liver accounted for most of the fructose clearance by the gut. contribute to observations that portal fructose levels peaked

Cell Metabolism 33, December 7, 2021 2335


ll
Review

Figure 2. Fructolysis and associated biochemistry


Fructose is transported into enterocytes and hepatocytes via GLUT5 and GLUT2, respectively. Upon entering the cells, fructose is phosphorylated by KHK to
F1P. Energy depletion resulting from robust fructose phosphorylation leads to activation of AMPD2 and uric acid production. F1P is cleaved by ALDOB to DHAP
and GA. GA is phosphorylated by triose-kinase (TKFC) to GA3P. Both DHAP and GA3P mix with triose-phosphates common to the glycolytic and gluconeogenic
carbon pools. In hepatocytes, F1P allosterically inhibits PGYL to enhance glycogen synthesis and disrupts the interaction between GCK and GCKR, allowing
GCK to translocate from the nucleus to the cytoplasm and catalyze phosphorylation of glucose, further increasing the hexose- and triose-phosphate carbon
pools. Fructose-derived substrate has numerous fates, including use in de novo lipogenesis (DNL), both through direct and indirect pathways via microbiome-
derived acetate. KHK, ketohexokinase; AMPD3, adenosine deaminase; IMP, inosine monophosphate; ALDOB, aldolase B; TKFC, triokinase and FMN cyclase;
GA, glyceraldehyde; DHAP, dihydroxyacetone phosphate; GA3P, glyceraldehyde 3-phosphate; PYGL, glycogen phosphorylase L; GYS2, glycogen synthase 2;
PKLR, pyruvate kinase, liver and red blood cell; PEP, phosphoenolpyruvate.

at 0.3 mM in mice after fructose gavage, whereas peak portal finity for fructose, has been reported to contribute to fructose
fructose concentrations were 3- to 5-fold higher in other animals, transport in both enterocytes and hepatocytes (DeBosch et al.,
including humans (Crossley and Macdonald, 1970; Holdsworth 2012, 2014; Schmidt et al., 2009). However, it predominately lo-
and Dawson, 1965; Topping and Mayes, 1971). Additional calizes to endosomal and lysosomal compartments rather than
detailed studies are required to clarify the importance of intesti- the cell surface (Alexander et al., 2020; Schmidt et al., 2009).
nal versus liver fructose metabolism in larger animals, including Upon transport into the cell, fructose is metabolized by a
humans. cascade of three ‘‘fructolytic’’ enzymes, originally delineated
Nevertheless, these studies led to a novel and important by Henri Hers during his graduate studies (Hers, 1957) (Figure 2).
concept that intestinal fructose metabolism might shield the liver These fructolytic enzymes efficiently convert the hexose fructose
from excessive exposure to dietary fructose, where its meta- into triose-phosphates, which are incorporated into the cellular
bolism may be particularly deleterious (Jang et al., 2018). This pools of triose-phosphates generated via glycolysis and gluco-
issue will be explored in more detail later in this review. neogenesis. The first step in fructolysis is the rapid and irrevers-
ible phosphorylation of fructose to fructose-1-phosphate (F1P)
CELLULAR AND INTERMEDIARY FRUCTOSE catalyzed by ketohexokinase (KHK, also known as fructokinase).
METABOLISM This metabolite is specific to the fructolytic pathway and is not
shared with glycolysis or gluconeogenesis. Compared with
Fructose transport across cell membranes relies on facilitative glucose, fructose is a poor substrate for hexokinases I, II, and
hexose carriers, such as GLUT5, in enterocytes, as there is no III (Km glucose < 0.1 mM; Km fructose 1–10 mM) (Diggle
known active transporter. In the liver, the other major site of fruc- et al., 2009; Grossbard and Schimke, 1966; Middleton, 1990).
tose metabolism, GLUT5 is not expressed at high levels, and At typical circulating concentrations, glucose potently inhibits
GLUT2 is likely the major membrane transporter for both glucose fructose metabolism by these hexokinases (Diggle et al., 2009;
and fructose. GLUT8 (also known as Slc2a8), which also has af- Froesch and Ginsberg, 1962). This is particularly true for

2336 Cell Metabolism 33, December 7, 2021


ll
Review

glucokinase (GCK), the predominate hexokinase in hepatocytes patic metabolism of glucose and fructose is not entirely indepen-
and pancreatic beta cells (Km for glucose 5 mM, Km fructose > dent of each other. F1P, the product of KHK-mediated fructose
100 mM) (Diggle et al., 2009; Froesch and Ginsberg, 1962; Mid- phosphorylation, potently relieves the inhibitory effect of GCKR
dleton, 1990; Pollard-Knight and Cornish-Bowden, 1982). on GK, permitting its translocation to the cytosol and enhancing
Unlike many other hexokinases, KHK is inhibited neither allo- hepatic glucose uptake and metabolism (Niculescu et al., 1997).
sterically by ATP nor by other signals of cellular energy suffi- Thus, the ‘‘catalytic’’ amounts of fructose can markedly enhance
ciency, nor by its immediate product. The capacity of KHK to hepatic glucose (Petersen et al., 2001; Van Schaftingen, 1994).
phosphorylate fructose is comparable with the capacity of F1P may also activate pyruvate kinase, the terminal step in
GCK to phosphorylate glucose in rodent livers and 10-fold higher glycolysis (Eggleston and Woods, 1970), and can inhibit
in human livers (Heinz et al., 1968). Moreover, whereas KHK is glycogen phosphorylase (Kaufmann and Froesch, 1973; Thur-
constitutively active, hepatic GCK is sequestered in the nucleus ston et al., 1974; Van Den Berghe et al., 1973). Through these
in an inhibited state by glucokinase regulatory protein (GCKR) F1P-mediated activities, the fructose contained in dietary sugars
(Agius, 2008; Niculescu et al., 1997). The sequestration and inhi- as part of a meal may act as a signal to modulate hepatic fuel
bition of GCK by GCKR and the stabilization of this inhibited state metabolism.
by the glycolytic intermediate fructose-6-phosphate (F6P) limit Recent work has firmly established additional roles for F1P as
the net hepatic glucose uptake and hepatic glucose clearance a key signaling molecule. Taylor et al. recently demonstrated that
after oral ingestion of an isolated glucose load (Agius, 2008; F1P can bind the M2 isoform of pyruvate kinase (PKM2) and limit
Pollard-Knight and Cornish-Bowden, 1982; Van Schaftingen, the formation of highly active PKM2 tetramers (Taylor et al.,
1994). In contrast, because of the low Km and high activity of 2021). PKM2 monomers and dimers are capable of binding
KHK for fructose and the lack of regulation of its activity, the ma- and transactivating HIF-1a, a transcription factor essential for
jority of fructose that arrives at the liver via the portal circulation is cellular adaptation to hypoxia (Luo et al., 2011). This protection
readily extracted and little reaches the systemic circulation. against hypoxia increases enterocyte cell survival, leads to elon-
The consumption of ATP by KHK-mediated fructose phos- gation of the intestinal villus, and increases absorptive capacity.
phorylation can be so robust that intravenous infusions of large Thus, fructose feeding, by increasing the intestinal absorptive
fructose loads cause an acute decline in the hepatocellular surface, may enhance energy harvest and potentially contribute
ATP-to-AMP ratio and a marked decline in free phosphate as it to obesity in calorie-rich environments.
is sequestered in F1P (Bode et al., 1973; Mayes, 1993). F1P Following fructose phosphorylation, F1P is cleaved by
can rapidly accumulate to millimolar concentrations in hepato- aldolase B (ALDOB) to dihydroxyacetone phosphate (DHAP)
cytes (Woods et al., 1970). These changes in F1P levels and and glyceraldehyde. Glyceraldehyde is then phosphorylated by
cellular energy status are detectable by non-invasive 31P mag- a triokinase (TKFC, triokinase and FMN cyclase) to glyceralde-
netic resonance spectroscopy in human livers (Oberhaensli hyde 3-phosphate (G3P). DHAP and G3P then enter the glyco-
et al., 1986). Following fructose infusion, hexose-phosphate lytic/gluconeogenic carbon pools.
levels increased 7-fold within minutes and steadily declined to Although cellular and organismal fuel status does not regulate
basal levels after 20 min. Hepatic ATP levels declined by fructolytic flux, fuel status does impact the fate of fructose-
more than 80% within minutes and recovered to only 50% of derived triose-phosphates that enter the central carbon pools
basal levels after 1 h (Oberhaensli et al., 1986). Whereas intrave- (Mayes, 1993). For instance, in fasted or starved animals, when
nous infusions of fructose can cause acute and profound PFK activity is limited by decreased fructose 2,6-bisphosphate
changes in hepatic metabolite levels and energy status, oral fruc- levels, fructose-derived triose-phosphates will be routed toward
tose administration results in more modest changes, likely due to glucose production (Exton and Park, 1967; Exton et al., 1966). In
the substantial metabolism of fructose within the intestine (Nie- contrast, in fed animals, fructose-derived triose-phosphates
woehner et al., 1984). Nevertheless, a recent NMR study indi- may be preferentially metabolized to pyruvate and released as
cates that ATP depletion can be detected in the human liver lactate or used as substrate for lipogenesis (Topping and Mayes,
following an oral 75 g fructose challenge, indicating that hepatic 1976). Nutrients co-ingested with fructose may also affect its
fructose metabolism in humans is substantial (Bawden metabolic fate. For instance, insulin that is secreted when
et al., 2016). glucose and fructose are ingested together may be important
Phosphate potently inhibits the enzyme AMP deaminase for the full effects of fructose to stimulate glycogen synthesis
(Ampd2), which catalyzes the rate-limiting step in purine degra- (Topping and Mayes, 1976). Thus, the fate and effects of in-
dation and uric acid production (van den Berghe et al., 1977). gested fructose on systemic fuel homeostasis depend on sys-
Thus, the decline in free phosphate resulting from fructose phos- temic fuel status.
phorylation potently activates Ampd and uric acid production.
Fructose may also stimulate purine synthesis (Raivio and KETOHEXOKINASE: GENETIC LESSONS IN HUMANS
Becker, 1975). The crystallization of uric acid in joints causes AND MICE AND THERAPEUTIC IMPLICATIONS
the painful rheumatological condition of gout. Thus, fructose-
induced uric acid production may contribute to the association KHK catalyzes the committed step in fructose metabolism,
between SSB consumption and a risk of gout, but it also likely phosphorylating fructose to F1P. Loss-of-function KHK muta-
underlies the association between hyperuricemia and cardiome- tions in humans produce the condition called benign essential
tabolic disease (Jamnik et al., 2016; Yoo et al., 2005). fructosuria (Asipu et al., 2003; Steinmann et al., 2019). A large
Although the committed phosphorylation steps in glucose and and persistent increase in circulating fructose is produced
fructose metabolism are catalyzed by distinct enzymes, the he- when people with this condition consume foods containing

Cell Metabolism 33, December 7, 2021 2337


ll
Review

fructose (Laron, 1961; Steinmann et al., 1975). As the name of the bolism have also been implicated in other diseases, including
condition suggests, in people with this condition, 20% of the hepatocellular carcinoma and heart failure (Li et al., 2016; Mirt-
fructose is excreted in the urine following an oral or intravenous schink et al., 2015).
fructose load (Laron, 1961; Steinmann et al., 1975). The Together, these studies indicate that fructose metabolism in a
remainder is presumably phosphorylated and cleared by canon- tissue that expresses the KHKc isoform is critical for fructose-
ical hexokinases, possibly in the adipose tissue and the skeletal induced metabolic disease. As previously noted, KHKc is ex-
muscle (Froesch and Ginsberg, 1962). Prior to the widespread pressed in both the intestine and the liver and the intestine is
use of glucose oxidase in assays to measure glucose levels in capable of substantial fructose metabolism. Knockout of intesti-
blood and urine samples, cases of benign essential fructosuria nal KHK exacerbates fructose-induced metabolic disease,
came to clinical attention when the use of Benedict’s reagent whereas liver-specific knockout or knockdown of KHK protects
incidentally detected high levels of reducing sugars in the urine, against fructose-induced metabolic disease (Andres-Hernando
and glucose was subsequently excluded with more specific et al., 2020b; Jang et al., 2020; Softic et al., 2018). These results
testing. Because Benedict’s reagent is no longer in regular support the hypothesis that intestinal fructose metabolism may
use, and because there are no known or reported adverse shield the liver from fructose exposure, where its metabolism
clinical consequences associated with loss of KHK activity, indi- may be particularly harmful. Mechanisms by which hepatic
viduals with this condition likely go undetected. Although this metabolism may contribute to metabolic disease will be ad-
condition appears to be uncommon, review of data from exome dressed in the subsequent sections of this review.
sequencing projects indicates that there is no strong selection Given the evidence that fructose metabolism via KHK is
against missense or loss-of-function variants in the KHK gene required for fructose-induced metabolic disease and that inacti-
in human populations (Johnston et al., 2021). vating mutations in KHK are not associated with adverse clinical
Despite increased circulating fructose in people with benign effects in humans, targeting the inhibition of KHK as a therapeu-
essential fructosuria, there is no increase in blood insulin and tic strategy seems to hold great promise. Recently, pharmaceu-
there are no reports of diabetic complications or increases in tical companies have initiated programs aimed at developing
%HbA1c (Petersen et al., 1992; Steinmann et al., 1975). The novel potent and specific KHK inhibitors, which are beginning
absence of pathology indicates that fructose metabolism via to provide data in pre-clinical and early-phase clinical trials (Fu-
KHK is essential for fructose-induced metabolic disease. This tatsugi et al., 2020; Gutierrez et al., 2021; Kazierad et al., 2021).
hypothesis is further supported by a mouse with global inactiva- Pre-clinical studies in rats demonstrated that the KHK inhibitor
tion of KHK (Ishimoto et al., 2012). Although KHK knockout mice PF-06835919 dose-dependently attenuated hyperinsulinemia,
showed a decreased preference for fructose-supplemented hypertriglyceridemia, and steatosis provoked by a high-fructose
drinking water compared with controls, when matched for total diet. These protective effects were associated with marked in-
energy and fructose intake, KHK knockout mice were fully pro- creases in circulating fructose and fructosuria and were accom-
tected from fructose-induced increases in body weight, fat panied by an attenuated induction of hepatic ChREBP activity
mass, serum insulin, and hepatic steatosis (Ishimoto et al., and DNL (Gutierrez et al., 2021). Importantly, PF-06835919
2012). Thus, fructose metabolism via KHK is essential for fruc- also enhanced metabolic indices in rats on an ‘‘American diet,’’
tose-induced metabolic disease. indicating that KHK inhibition may be healthful outside of
KHK is expressed as two distinct isoforms from a single gene extreme fructose exposures. To this end, a randomized, dou-
(Hayward and Bonthron, 1998). Mutually exclusive splicing of ble-blind, placebo-controlled Phase 2a study (NCT03256526)
exons 3a and 3c in KHK pre-mRNA produces distinct mRNAs in adults with baseline NAFLD, PF-06835919, produced an
that encode KHKa and KHKc protein isoforms, respectively. 18.7% reduction in liver fat after 6 weeks of treatment with a
The Km of KHKc for fructose is 0.5 mM, whereas the Km of 300 mg but not a 75 mg daily dose (Kazierad et al., 2021). Treat-
KHKa for fructose is more than an order of magnitude higher ment with the KHK inhibitor resulted in increased fructosuria
(Diggle et al., 2009). KHKa is expressed at low levels ubiqui- consistent with robust KHK inhibition in vivo. This was also
tously. In contrast, KHKc is expressed at high levels selectively associated with improvements in inflammatory markers and in-
in key metabolic tissues, including the liver, small intestine, and creases in circulating adiponectin levels. The drug was well toler-
kidney (Diggle et al., 2009). In addition to knockout mice in which ated. This short-term study provides optimism that inhibiting
both KHK isoforms were globally ablated, Ishimoto and col- fructose metabolism may provide a new therapeutic strategy
leagues generated a mouse in which the KHKa isoform was for treating cardiometabolic disease soon.
selectively deleted. The selective ablation of KHKa exacerbated
fructose-induced disease, indicating that, despite its low levels, ADDITIONAL FRUCTOLYTIC ENZYMES: GENETIC
KHKa likely does contribute to systemic fructose metabolism LESSONS IN HUMANS AND MICE
(Ishimoto et al., 2012). In contrast, selective deletion of KHKc re-
vealed that it is primarily responsible for fructose-induced meta- Whereas the global KHK loss-of-function mutations are benign,
bolic disease. This conclusion is further supported by recent the loss-of-function mutations in Aldob, the second step in fruc-
work conducted by Nikolaou et al. on identifying the antagonistic tolysis, cause hereditary fructose intolerance (HFI), an autosomal
actions of RNA-binding proteins A1CF and hnRNPH1/2 that recessive Mendelian disorder with a significant morbidity (Ali
regulate KHKa/c splicing (Nikolaou et al., 2019). A1CF knockout et al., 1998; Hers and Joassin, 1961; Perheentupa et al., 1972).
reduces the expression of the KHKc isoform and protects mice Individuals with this condition tolerate carbohydrates containing
from fructose-induced metabolic disease (Nikolaou et al., glucose or galactose, whereas ingestion of sucrose or fructose
2019). Regulated splicing of KHK and altered fructose meta- triggers abdominal pain, nausea, and vomiting accompanied

2338 Cell Metabolism 33, December 7, 2021


ll
Review

by acute hypoglycemia. This condition often presents in infants part, due to signaling effects of increased F1P generated from
as a failure to thrive accompanied by liver toxicity after weaning endogenously produced fructose (Oppelt et al., 2015). The in-
from breast milk or the introduction of a fructose-containing for- crease in steatosis in Aldob knockout mice can be reversed by
mula. Aversion to sweet-tasting foods is also common (Cham- inhibiting KHK, which further supports the hypothesis that F1P
bers and Pratt, 1956). Persistent exposure to fructose ultimately is a critical signaling molecule in the pathogenesis of this condi-
leads to liver and kidney failure and, ultimately, death (Baerlocher tion (Lanaspa et al., 2018a).
et al., 1978). Similar morbidity and mortality can also be pro- As the adverse effects of HFI depend on fructolytic flux, the in-
voked by administration of sorbitol, a metabolite that can be hibition of KHK could potentially be therapeutic for HFI as it is for
used to produce fructose endogenously via the polyol pathway other forms of fructose-induced disease. In support of this
(Ali et al., 1998). People with this condition often learn to avoid strategy, Lanaspa and colleagues noted that Osthole, a plant-
all dietary sources of fructose, and are often devoid of dental derived coumarinic derivative with pleiotropic biological effects,
caries, supporting evidence that dietary sucrose is a major including KHK inhibition, is able to prevent adverse metabolic
contributor to poor dental health (Ali et al., 1998; Marthaler and symptoms in Aldob knockout mice (Lanaspa et al., 2018a; Le
Froesch, 1967; Touger-Decker and van Loveren, 2003). Elimina- et al., 2016). These results provide hope that KHK inhibitors
tion of all dietary sucrose, fructose, and sorbitol is the mainstay may be useful to markedly expand food options in people with
of treatment and improved health can be observed within days of this condition and warrant trials with newly developed potent
fructose restriction (Mock et al., 1983; Steinmann et al., 2019). and specific KHK inhibitors.
Properly treated, HFI patients may experience normal health, The Aldob cleavage reaction generates two metabolites from
growth, and lifespan. F1P—DHAP and glyceraldehyde. DHAP requires no further
Lean and healthy adults with HFI on low-fructose diets have metabolism to join the glycolytic triose-phosphate pool. Howev-
increased intrahepatic triglyceride and glucose intolerance (Si- er, glyceraldehyde, a potentially reactive and toxic metabolite, is
mons et al., 2019). This speaks to the importance of F1P as a phosphorylated by Tkfc to the glycolytic intermediate glyceral-
metabolic signal and indicates that increased fructolytic flux dehyde-3-phosphate, completing the final step in fructolysis.
per se is not necessary for the pathogenesis of these features. Tkfc has been studied less intensively than the preceding two
Additional support for the importance of F1P as a signaling mole- steps, in part, because there are no known clinical conditions
cule comes from recent studies in adults heterozygous for HFI associated with mutations in this enzyme. Mice with liver-spe-
(hHFI) and controls (Debray et al., 2021). Whereas a high-fruc- cific or global Tkfc knockdown demonstrate liver inflammation
tose diet increased fasting uric acid and insulin resistance simi- and evidence of fructose malabsorption when challenged with
larly in both hHFI patients and controls, in the post-prandial high-fructose diets, potentially related to toxicity of fructose-
condition, plasma uric acid and liver insulin resistance increased derived glyceraldehyde (Liu et al., 2020; Sillero et al., 1969).
only in the hHFI patients (Debray et al., 2021). These mice demonstrated increased glycerate and serine syn-
Fructose toxicity in HFI is attributed, in part, to the deleterious thesis, suggesting metabolism of glyceraldehyde via aldehyde
effect of the marked and prolonged accumulation of F1P in he- dehydrogenase (Sillero et al., 1969). The recent observation in
patocytes, enterocytes, and epithelial cells of the proximal tu- mice that a substantial fraction of ingested fructose is exported
bule (Oberhaensli et al., 1987; Steinmann et al., 2019; Van Den from enterocytes into portal blood as glycerate suggests that
Berghe et al. 1973). ATP depletion and uric acid production are Tkfc activity may be limiting for fructolysis in some conditions
also posited to contribute to toxicity in the affected cell types (Jang et al., 2018).
(Steinmann et al., 2019; Van Den Berghe et al. 1973). Hypoglyce-
mia that occurs acutely after fructose consumption in HFI may be MOLECULAR MEDIATORS OF FRUCTOSE ON LIPID
due to the effects of F1P glucokinase activation and also the ef- HOMEOSTASIS
fects of very high levels of F1P on inhibiting glycogenolysis and
gluconeogenesis (Rambaud et al., 1973; Steinmann et al., As previously noted, the potential deleterious effects of fructose
2019; Van Den Berghe et al. 1973). This may be mediated by on metabolism were initially recognized through its adverse ef-
an effect of F1P to inhibit glycogen phosphorylase (Van Den fects on lipid homeostasis. This included the ability of extremely
Berghe et al., 1973). high fructose exposures far exceeding those of typical Western
Although fructose metabolism in the liver, intestine, and kidney diets to provoke fasting hypertriglyceridemia and steatosis and
is greatly impaired in HFI, less than 20% of the fructose load is to enhance prandial lipemia within a matter of days. These
recovered in the urine in these patients, indicating that the major- changes in liver and circulating triglyceride levels are consis-
ity of the fructose is metabolized in other tissues (Landau et al., tently associated with increased hepatic DNL—the biochemical
1971). Adipose tissue metabolism of fructose to fructose-6- process by which new fatty acids are synthesized from precursor
phosphate by hexokinase II may be the major mode of fructose molecules (Geidl-Flueck et al., 2021; Lê et al., 2009; Wakil et al.,
metabolism in this condition, as may also be the case in benign 1983). Indeed, fructose ingestion can increase the amount of
essential fructosuria (Froesch and Ginsberg, 1962). newly synthesized fatty acids found within circulating triglyceride
The molecular physiology of HFI has been confirmed in a in hours (Hudgins et al., 2011). Moreover, newly synthesized fatty
mouse model with global knockout of Aldob (Oppelt et al., acids are increasingly recognized as major contributors to he-
2015). This model recapitulates features observed in human patic fat accumulation associated with obesity and NAFLD (Don-
patients, including morbidity and mortality, when exposed to nelly et al., 2005; Lambert et al., 2014; Smith et al., 2020).
high-fructose diets. On a fructose-free diet, these mice develop Fructose may facilitate DNL through multiple mechanisms
steatosis, as is the case for humans with HFI. This may be, in (Figure 3). Fructose-derived metabolites, such as F1P, act as

Cell Metabolism 33, December 7, 2021 2339


ll
Review

partner, GCKR (Agius, 2008; Van Schaftingen, 1994). This serves


to increase glycolytic flux and increases the concentration of
hexose- and triose-phosphates within the hepatocyte (Kim
et al., 2016). One or more of these glucose metabolites activates
ChREBP, leading to coordinate upregulation of the full comple-
ment of enzymes required for DNL (Dentin et al., 2004; Iizuka
et al., 2004; Katz et al., 2021; Kim et al., 2016). Increased glyco-
lytic flux via activation of GCK also provides additional substrate
and reducing equivalents, supporting both DNL and glycogen
synthesis.
ChREBP is an evolutionarily conserved, carbohydrate-sensing
transcription factor that is highly expressed in key metabolic cell
types, including enterocytes, hepatocytes, adipocytes, pancre-
atic beta cells, and proximal tubule cells of the kidney (Iizuka
et al., 2004; Katz et al., 2021; Uyeda and Repa, 2006). ChREBP
is required for carbohydrate-mediated induction of glycolytic
and lipogenic enzymes in the liver (Iizuka et al., 2004). It also
transactivates the expression of the complement of fructolytic
enzymes (Iizuka et al., 2004; Ma et al., 2006). It was discovered
based upon its ability to mediate upregulation of the glycolytic
enzyme liver pyruvate kinase (Pklr) independent of insulin
signaling in primary hepatocytes exposed to high-glucose cul-
ture media (Yamashita et al., 2001). However, in vivo, fructose,
but not glucose, gavage acutely and robustly upregulates the
expression of hepatic ChREBP-beta, a potent ChREBP isoform,
along with its fructolytic, glycolytic, and lipogenic targets (Kim
et al., 2016). Liver-specific ChREBP knockdown or knockout
prevents sucrose- and fructose-mediated induction of DNL en-
zymes and DNL activity (Erion et al., 2013; Kim et al., 2017;
Linden et al., 2018).
In humans, genetic variants in the ChREBP locus are strongly
associated with increased circulating triglyceride levels and
reduced HDL-cholesterol levels, the dyslipidemia characteristi-
cally associated with the metabolic syndrome (Kooner et al.,
2008). Fructose consumption acutely increases ChREBP activ-
Figure 3. Fructose activates metabolic gene expression
ity, which has been observed in livers of humans with obesity,
Fructose-derived metabolites act as signaling molecules to activate metabolic
transcriptional programs. F1P derived from fructose metabolism activates diabetes, and fatty liver disease (Eissing et al., 2013; Kursawe
GCK, and the combination of fructose- and glucose-derived hexose phos- et al., 2013). This suggests that ChREBP links hepatocellular
phates activates ChREBP, which coordinately regulates enzymes involved in hexose availability to both DNL, as well as very-low-density lipo-
fructolysis, glycolysis, glucose production, lipogenesis, and VLDL packaging
and export. Fructose-derived metabolites may also activate SREBP1c. Both protein (VLDL) packaging and export and may mediate the DNL
ChREBP and SREBP1c are coactivated by PGC1b to further coordinate and hypertriglyceridemia after fructose feeding in individuals
and enhance transcription of metabolic programs. ACLY, ATP citrate lyase; with obesity and insulin resistance. Recent evidence demon-
ACACA, acetyl-CoA carboxylase a; FASN, fatty acid synthase; GPAT, glycerol-
3-phosphate acyltransferases; Agpat, acylglycerol-3-phosphate acyltransfer-
strates that genetic variants in the ChREBP locus interact with
ase; DGAT, diacylglycerol acyltransferase; MTTP, microsomal triglyceride sugar consumption to impact circulating lipids in human popula-
transfer protein; TM6sf2, transmembrane 6 superfamily, member 2; TAG, tri- tions, further supporting a role for ChREBP as a central mediator
acylglycerol; VLDL, very-low-density lipoprotein. of sugar’s effects on cardiometabolic lipid risk factors in people
(Haslam et al., 2021). ChREBP may serve a similar function to
signaling molecules to promote DNL (Kim et al., 2016). As noted promote DNL and chylomicron packaging in enterocytes after
earlier, this is supported by evidence in patients and animal fructose feeding (Haidari et al., 2002).
models of HFI, where increased F1P but decreased fructolytic Liver ChREBP knockdown in fructose-fed rats reduced circu-
flux are associated with increased steatosis. Nevertheless, lating triglyceride levels, confirming a role for ChREBP in fruc-
increased fructolytic flux may also provide substrate and tose-mediated dyslipidemia (Erion et al., 2013). The effects of
reducing equivalents to support fatty acid synthesis (Wakil ChREBP on circulating triglycerides are likely mediated through
et al., 1983). Additionally, uric acid generated as a result of rapid its regulation of key enzymes involved in VLDL packaging and
fructose catabolism has been suggested to promote DNL (La- export, including microsomal triglyceride transfer protein (Mttp)
naspa et al., 2012). and transmembrane 6 superfamily, member 2 (Tm6sf2) (Lei
The predominant mechanism by which fructose induces DNL et al., 2020; Niwa et al., 2018). Hepatic ChREBP may also limit
in the liver is through its effects on increasing hepatocellular F1P, circulating triglyceride clearance through its ability to stimulate
which causes the dissociation of GCK from its inhibitory binding production of proteins, such as apolipoprotein C3 (Apoc3) and

2340 Cell Metabolism 33, December 7, 2021


ll
Review

angiopoietin-like 8 (Angptl8), which inhibit lipoprotein lipase (Lpl) gai et al., 2009). As a result, PGC-1b appears to promote multiple
activity and the clearance of circulating triglycerides (Caron adverse sugar-mediated phenotypes and poses interesting po-
et al., 2011; Fu et al., 2014). Although liver-specific ChREBP tential as a therapeutic target.
knockout or knockdown fully abrogates DNL, under some condi- Recent work has recast the role of fructose as a lipogenic sub-
tions, this has little or no effect on steatosis (Erion et al., 2013; strate within hepatocytes. The path by which fructose-derived
Kim et al., 2017; Linden et al., 2018). Linden et al. suggested intermediates can be used as substrates for fatty acid synthesis
that the reduction in DNL in ChREBP KO mice is balanced by a within hepatocytes requires synthesis of citrate in the mitochon-
reduction in VLDL packaging export, leading to little net change dria, which is then exported from the mitochondria and cleaved
in liver fat content (Linden et al., 2018). Additionally, although by ATP citrate lyase (Acly) to generate cytosolic acetyl-CoA, the
DNL is a significant contributor to steatosis, particularly in pa- primary precursor for new fat synthesis (Sul et al., 1984). Con-
tients with obesity and NAFLD, the majority of lipid in steatotic trary to conventional wisdom, Zhao and colleagues showed
livers originates from circulating free fatty acids rather than that fructose-derived carbons could be incorporated into the
DNL (Donnelly et al., 2005; Smith et al., 2020). These results high- newly synthesized fatty acids in the liver in Acly knockout
light that DNL and steatosis are regulated by distinct processes mice, demonstrating that fructose-derived citrate is not required
and can be dissociated from each other. The importance of DNL for its use in DNL (Zhao et al., 2020). In this study, fructose car-
to both NAFLD and the dyslipidemia of metabolic syndrome re- bons were used as lipogenic substrate via an indirect path. Fruc-
quires further investigation. tose catabolized by microbiota in the distal gut to acetate was
ChREBP does not act alone in regulating the expression of absorbed, transported via the portal vein, taken up by hepato-
DNL enzymes and other mediators of lipid homeostasis. The ste- cytes, and converted to acetyl-CoA by an acetyl-CoA synthetase
rol regulatory element-binding protein 1c (SREBP1c) is another (Acss2). When fructose was consumed more gradually, both the
master regulator of lipogenic enzymes that responds to both nu- conventional, direct citrate pathway and the microbiota-acetate
trients and hormones (Shimano and Sato, 2017). The effects of pathway appeared to contribute fructose carbons to the newly
ChREBP and SREBP1c are synergistic on lipogenic targets synthesized fatty acids (Zhao et al., 2020).
(Linden et al., 2018). Insulin strongly promotes SREBP1c activity Fructose-derived metabolites may provide signals that regu-
via a pathway that is dependent on AKT and mechanistic target late additional aspects of intermediary metabolism. For instance,
of rapamycin complex 1 (mTORC1) (Shimano and Sato, 2017). the second step in DNL is mediated by acetyl-CoA carboxylase
Although fructose does not directly stimulate insulin secretion, (ACACA), which catalyzes the carboxylation of acetyl-CoA, the
increased adiposity and hyperinsulinemia associated with rate-limiting step in DNL, to generate malonyl-CoA (López-Casil-
chronic consumption of fructose might enhance SREBP1c ac- las et al., 1988). Malonyl-CoA serves as an important signaling
tivity. molecule, as it potently inhibits carnitine palmitoyltransferase
Along with activation of ChREBP, fructose feeding can activate 1A (Cpt1a), an enzyme required for translocation of fatty acids
SREBP1c independently of hepatic insulin signaling in liver-spe- into the mitochondria (McGarry, 2002). Thus, fructose-derived
cific insulin receptor-knockout mice (Haas et al., 2012). Recent signals also serve to inhibit fatty acid oxidation (Topping and
work indicates that DHAP, a product of both glycolysis and fruc- Mayes, 1972). Recent evidence suggests that fructose-derived
tolysis, provides a signal to activate mTORC1, which is important metabolites might also be used to post-translationally modify
in activating SREBP1c (Orozco et al., 2020). However, fructose mitochondrial proteins, including Cpt1a, which may also
feeding appears to inhibit, rather than activate, mTOR (Hu contribute to the inhibition of fatty acids (Softic et al., 2019). In-
et al., 2018). The mechanisms mediating fructose-induced acti- hibiting fatty acid oxidation is another mechanism by which fruc-
vation of SREBP1c remain uncertain. Additional recent evidence tose-derived signals may contribute to the development of
in humans supports the hypothesis that increased lipogenic steatosis.
enzyme expression and lipogenesis in NAFLD patients is depen-
dent on substrate-derived signals, rather than systemic hormonal FRUCTOSE AND NAFLD
cues, activating both ChREBP and SREPB1c (Ter Horst
et al., 2020). Hepatic steatosis constitutes the earliest stage of NAFLD and is
Some, but not all, studies suggest that fructose consumption diagnosed when more than 5% of liver cells contain lipid droplets
promotes ER stress, which may also induce proteolytic cleavage on biopsy or, more commonly, by specific estimates of liver fat
and activation of SREPB1c (Flamment et al., 2012; Kammoun content based on imaging. In 10%–15% of individuals, steatosis
et al., 2009; Kim et al., 2017; Marek et al., 2015). ER stress can progresses to non-alcoholic steatophepatitis (NASH), which is
also activate the transcription factor XBP1, which may enhance characterized by inflammation and fibrosis and, in turn, may
the expression of lipogenic enzymes independently of other lipo- evolve further to cirrhosis and increased risk for hepatocellular
genic transcription factors (Lee et al., 2008; Marek et al., 2015). carcinoma. In individuals with NAFLD, steatosis strongly associ-
PPAR-g coactivator 1b (PGC-1b) is a transcriptional coactiva- ates with increased DNL and variably associates with changes in
tor that can enhance transcriptional activity of both ChREBP and VLDL secretion and fat oxidation (reviewed in Ter Horst and Ser-
SPREBP1c, as well as other important metabolic transcription lie, 2017). Because of fructose’s ability to promote DNL and
factors, including PPAR-g, PPAR-a, estrogen-related receptors inhibit fat oxidation, investigators have posited that fructose con-
(ERRs), and liver X receptor (LXR) (Chambers et al., 2013; Lin sumption is a major driver of NAFLD (Jensen et al., 2018; Lim
et al., 2002, 2005a, 2005b). In fructose-fed rats, the knockdown et al., 2010). Although studies administering very high doses of
of hepatic PGC-1b prevented increases in adiposity, lipogen- fructose support the possibility that fructose promotes steatosis
esis, steatosis, hyperinsulinemia, and hypertriglyceridemia (Na- in humans, so far, evidence that fructose when consumed at

Cell Metabolism 33, December 7, 2021 2341


ll
Review

levels typical of Western diets is associated with the develop- after administration of fructose, recovery of energy status is de-
ment of NAFLD is mixed. This pertains to both prospective layed in the livers of patients with NASH (Cortez-Pinto et al.,
epidemiological cohorts evaluating sugar and fructose con- 1999). However, follow-up studies in participants with obesity
sumption, as well as short-term intervention studies comparing showed mixed results (Bawden et al., 2016; Nair et al., 2003).
the addition of isocaloric fructose versus glucose (Agebratt These human studies are complemented by studies performed
et al., 2016; Chiu et al., 2014; Chung et al., 2014; Kanerva in fructose-fed mice (Lanaspa et al., 2012; Softic et al., 2019).
et al., 2014; Ma et al., 2015; Ouyang et al., 2008; Schwarz Fructose feeding produced evidence of altered mitochondrial
et al., 2015). Two fructose restriction studies support the possi- morphology, an altered mitochondrial proteome, and increased
bility that reducing fructose consumption, even isocalorically, re- reactive oxygen species that could potentially adversely impact
duces DNL and liver fat in humans, but these studies lacked con- mitochondrial function and contribute to changes in energy sta-
trol arms, which limit definitive conclusions (Schwarz et al., 2017; tus and liver disease (Lanaspa et al., 2012; Softic et al., 2019).
Volynets et al., 2013). Recently, a small, well-controlled sugar re- Recent investigations have focused not only on the direct ef-
striction study was performed in adolescent boys with histolog- fects of fructose metabolism within the liver but also indirect ef-
ical evidence of NAFLD (Schwimmer et al., 2019). Free sugars fects potentially mediated by changes in the microbiome and gut
were restricted to less than 3% of daily calories for 8 weeks barrier function. Complex associations between changes in the
and this was compared with ‘‘regular’’ diet controls. Sugar re- gut microbiome and NAFLD are of increasing interest (Sharpton
striction was associated with a greater reduction in liver fat, as et al., 2021). Unabsorbed fructose serves as fuel for the intestinal
well as a greater reduction in alanine aminotransferase, a circu- microbiota and may contribute to a wide range of changes in in-
lating marker of liver injury. In the last year, another small, well- testinal microbial communities (reviewed in Lambertz et al.,
controlled restriction study was completed in overweight adults 2017). Fructose-dependent alterations in the intestinal micro-
(Simons et al., 2021). Dietary fructose was restricted to less than biota or direct effects on the epithelial cells may disrupt barrier
10 g per day, and individuals were randomly assigned to supple- function, leading to the progression of NAFLD due to increased
mentation with fructose to achieve fructose intake similar to intestinal permeability (Bergheim et al., 2008). This may increase
baseline (control group) or isocaloric glucose over 6 weeks (re- TNF-alpha and endotoxin in the portal blood, effects that can be
striction group). The fructose-restricted group achieved small, attenuated with antibiotics (Bergheim et al., 2008; Todoric et al.,
but statistically significant, reductions in liver fat compared 2020). Recent work suggests that prolonged exposure to 30%
with controls (Simons et al., 2021). Perhaps the most compelling fructose in drinking water causes ER stress within colonic enter-
evidence that fructose is a common contributor to NAFLD de- ocytes, leading to barrier breakdown and endotoxemia in mice
rives from a clinical study in which participants with steatosis (Todoric et al., 2020). This endotoxemia induces TNF-alpha in
were administrated a KHK inhibitor, thus blocking fructose meta- liver myeloid cells, which can induce ER stress in hepatocytes,
bolism in the liver, intestines, and kidneys (Kazierad et al., 2021). leading to cleavage and activation of SREBP1 (Kim et al.,
Participants with a mean baseline liver fat of 15% were asked to 2018; Todoric et al., 2020). However, in normal physiological cir-
maintain their normal diets without fructose supplementation for cumstances, little ingested fructose reaches the colon. Indeed,
6 weeks, during which time they were treated with either one of 2 weeks of fructose supplementation in a small preliminary pro-
two doses of KHK inhibitor or placebo. At the end of the study, spective cohort of adults with obesity did not show any changes
study participants who took the higher dose of KHK inhibitor in microbiome, fecal metabolites, or indices of gut permeability,
demonstrated a greater than 18% reduction in liver fat compared inflammation, or endotoxemia (Aleman et al., 2020).
with placebo. These data strongly support the proposal that
quantities of fructose typical of a modern Western diet contribute FRUCTOSE EFFECTS ON GLUCOSE HOMEOSTASIS
significantly to the increasing prevalence of NAFLD.
A minority of patients with NAFLD progress to more advanced Along with changes in liver fat and circulating triglycerides, short-
forms of liver injury involving inflammation and fibrosis. Some ev- term hypercaloric sucrose or fructose feeding can rapidly induce
idence suggests that increased fructose consumption may pro- fasting hyperinsulinemia in humans and animal models (Lê et al.,
mote progression to more advanced forms of NAFLD in children, 2009; Reaven, 1988; Ter Horst et al., 2016). This occurs despite
adolescents, and adults (Abdelmalek et al., 2010; Mosca et al., the fact that fructose in isolation does not stimulate insulin secre-
2017). The ability of fructose to promote progression could be tion from pancreatic beta cells (Curry, 1989). Presumably, this
due to its effects on enhancing lipogenesis and steatosis, but hyperinsulinemia is an indicator of liver and/or peripheral insulin
other mechanisms related to unique aspects of its metabolism resistance. In rats, the dose and duration of sugar exposure have
could also contribute. Emerging data suggest that hepatocellular significant effects on which tissues are affected. Compared with
energy homeostasis is abnormal in NAFLD, and these energetic glucose, very high doses of fructose and sucrose (>60% of total
derangements may be exacerbated by fructose. As previously energy) can induce both hepatic and peripheral insulin resis-
noted, the rapid increase in hexose phosphates and the depletion tance within 2 months. However, more moderate doses (18%
of ATP that can result from fructose phosphorylation within hepa- of total energy) for twice as long elicited isolated hepatic insulin
tocytes can be detected non-invasively by 31P magnetic reso- resistance (Pagliassotti and Prach, 1995; Pagliassotti et al.,
nance spectroscopy (MRS). Using this and related techniques, 1994; Thresher et al., 2000). Similar to the case in rodents, isoca-
the majority of MRS studies detect alterations in energy status loric and moderate hypercaloric fructose feeding in humans has
in the livers of patients with NAFLD (Abrigo et al., 2014; Bawden preferential effects to induce hepatic over peripheral insulin
et al., 2016; Cortez-Pinto et al., 1999; Karczmar et al., 1989; Nair resistance (Aeberli et al., 2013; Schwarz et al., 2015; Ter Horst
et al., 2003; Traussnigg et al., 2017). A pilot study suggested that et al., 2016).

2342 Cell Metabolism 33, December 7, 2021


ll
Review

Pagliassotti and colleagues noted that gluconeogenic capac- sive energy intake, overweight, and obesity. Additional mecha-
ity was increased in primary hepatocytes isolated from high nisms independent of its hedonic value have also been invoked
sucrose and fructose feeding, possibly due to a fructose-depen- to explain why fructose might be particularly obesogenic. For
dent increase of glucose-6-phosphatase (G6pc), the final enzy- instance, fructose ingestion may have distinct effects on anorex-
matic step of glucose production (Bizeau et al., 2001; Wei igenic and orexigenic hormones, such as leptin and ghrelin, that
et al., 2004). ChREBP is essential for fructose-mediated upregu- impact feeding behaviors. Meal-associated increases in leptin
lation of both intestinal and hepatic G6pc (Kim et al., 2016; Ped- are diminished by fructose compared with glucose and fructose
ersen et al., 2007). The ChREBP-mediated increase of G6pc is may induce leptin resistance (Chotiwat et al., 2007; Shapiro
dominant over the antagonistic action of insulin to suppress it et al., 2008; Teff et al., 2004). Fructose ingestion less potently
(Kim et al., 2016). This ChREBP-G6pc signaling axis is suppresses ghrelin secretion compared with glucose ingestion
conserved in humans and likely contributes to hepatic insulin (Teff et al., 2004). All these effects could promote excessive
resistance in humans in the setting of high-sugar diets (Kim food intake and weight gain.
et al., 2016). Hepatic ChREBP knockdown in rats also attenuated Recent evidence derived from human genetics and non-clin-
peripheral insulin resistance by an unknown mechanism (Erion ical model organisms has suggested a negative feedback loop,
et al., 2013). Recent work demonstrated that knocking out whereby the consumption of fructose suppresses further sugar
KHK in mice prevented fructose-induced changes in adipose consumption. FGF21 is a liver-derived hormone that regulates
inflammation, insulin resistance, and a decline in circulating adi- systemic fuel homeostasis (Flippo and Potthoff, 2021). In hu-
ponectin (Marek et al., 2015). Mechanisms linking fructose meta- mans, non-human primates, and rodents, fructose consumption
bolism to changes in adipose function are uncertain. Fructose acutely and robustly increases hepatic production of FGF21 in a
may also cause hepatic insulin resistance by impairing hepatic ChREBP-dependent manner (Dushay et al., 2015; Iizuka et al.,
insulin signaling (Softic et al., 2020). Fructose-induced ER stress 2009; Kim et al., 2017; Talukdar et al., 2016). FGF21 signals via
and associated JNK kinase activation and signaling have been neural circuitry, including through glutamatergic neurons in the
invoked as a cause of hepatic insulin resistance (Sun et al., ventromedial hypothalamus, to suppress further carbohydrate
2015; Wei and Pagliassotti, 2004; Wei et al., 2005). As fructose intake (von Holstein-Rathlou et al., 2016; Jensen-Cody et al.,
feeding may cause steatosis and steatosis is commonly associ- 2020; Talukdar et al., 2016). Moreover, common genetic variants
ated with hepatic insulin resistance, one or more of the many in the FGF21 locus associate with sugar versus fat preferences in
putative lipid mediators implicated in lipid-mediated insulin human populations (Chu et al., 2013; Tanaka et al., 2013).
resistance may be involved (Petersen and Shulman, 2018). For The effect of fructose to suppress further carbohydrate con-
instance, fructose feeding has been associated with diacylgly- sumption may confound interventional trials that aim to study
cerol accumulation, PKC activation, and impaired insulin-medi- the metabolic effects of added sugars. Indeed, multiple rigorous
ated AKT2 activation (Jurczak et al., 2012; Nagai et al., 2009). studies have found that supplemental sugar markedly reduced
One other distinct mode of fructose-mediated hepatic insulin spontaneous consumption of other carbohydrates, limiting dif-
resistance has been described. Coate and colleagues noted ferences in total sugar consumption between experimental and
that hepatic glucose uptake was impaired in both high-fat and control participants (Ebbeling et al., 2020; Maersk et al., 2012).
high-fructose fed dogs (Coate et al., 2010, 2014). Whereas fat This strong, negative feedback loop must be considered when
impaired hepatic insulin action and AKT phosphorylation, fruc- designing and interpreting dietary intervention trials.
tose did not. Rather, fructose impaired hepatic Gck protein
expression, Gck activity, and hepatic glycogen synthesis. The FRUCTOSE CONTRIBUTIONS TO DYSREGULATED
mechanism mediating fructose’s negative effects on hepatic BLOOD PRESSURE
Gck is unknown.
Recent work indicates that fructose metabolism in the liver Whereas extreme fructose-enriched diets can induce hyperten-
may also regulate systemic amino acid metabolism. In particular, sion in rodents, the association between SSB consumption and
fructose-mediated activation of ChREBP upregulates and down- hypertension in humans is weaker, particularly when compared
regulates BCKDK and PPM1K, respectively (White et al., 2018). with other cardiometabolic risk factors (Hwang et al., 1987;
BCKDK is a kinase that phosphorylates and inhibits BCKDH, the Khan and Sievenpiper, 2016). Potential mechanisms by which
rate-limiting step in branched-chain amino acid (BCAA) oxida- fructose may affect blood pressure include effects on intestinal
tion (White and Newgard, 2019). This phosphorylation and inhibi- salt absorption, renal salt absorption and function, or endothelial
tion are reversed by the phosphatase PPM1K. As the liver is a function (reviewed in Klein and Kiat, 2015). Further, based on
major site of BCAA oxidation, this mechanism may contribute epidemiological associations and interventional studies in ani-
to the long-standing observation that increased circulating mal models, Johnson and colleagues have proposed that
branched-chain amino acids are associated with insulin resis- fructose-induced hyperuricemia is a major contributor to the
tance and impaired glucose homeostasis in humans (Newgard deleterious metabolic effects of fructose, including its adverse
et al., 2009; White and Newgard, 2019). effects on kidney function, blood pressure, and cardiovascular
risk (Ejaz et al., 2020). However, most but not all Mendelian
EFFECTS OF FRUCTOSE ON APPETITE, randomization studies indicate that genetic variants that causally
MACRONUTRIENT PREFERENCE, AND OBESITY affect uric acid levels and the risk of gout do not associate with
other adverse cardiometabolic traits (Ge et al., 2020; Jordan
The hedonic reward derived from consuming sugars containing et al., 2019; Tin et al., 2019; Yang et al., 2010). Moreover, recent
fructose contributes to its overconsumption, leading to exces- randomized, prospective interventional trials in humans aimed at

Cell Metabolism 33, December 7, 2021 2343


ll
Review

reducing uric acid production did not show benefits in patients fructose is produced endogenously from glucose, this may sug-
with type 1 diabetes or chronic kidney disease (Badve et al., gest that increased endogenous fructose production is a marker
2020; Doria et al., 2020). Thus, although uric acid production re- of worsening glycemia. However, even after adjustment for
mains an excellent marker of fructose metabolism, its causal role baseline glucose levels and SSB consumption, fasting blood
in fructose-associated pathologies remains controversial. fructose levels is associated with an increased risk of incident
Fructose feeding may promote hypertension by enhancing in- T2D in a dose-dependent manner, possibly due to a pathogenic
testinal salt absorption (Barone et al., 2009; Singh et al., 2008). ‘‘feedforward’’ loop whereby increased blood glucose stimu-
Fructose induces the intestinal anion exchanger Slc26a6 in a lates fructose production, which, in turn, exacerbates T2D and
GLUT5-dependent manner, and Slc26a6 knockout mice are cardiometabolic risk (Chen et al., 2020).
resistant to fructose-induced hypertension, but not to other The increased serum, urine, and tissue sorbitol concentrations
adverse metabolic effects of fructose feeding (Barone et al., present in individuals with diabetes have been implicated in the
2009; Singh et al., 2008). Fructose consumption suppressed development of diabetic microvascular complications (Brownlee,
renin expression consistent with salt overload, and reduced 2005; Lorenzi, 2007; Oates, 2002; Preston and Calle, 2010). The
renal salt excretion, which may further contribute to fructose- activity of polyol pathway may vary significantly within the popu-
induced hypertension. The effects of fructose on increased lation (Yoshii et al., 2001), raising the possibility that people with
blood pressure may be synergistic with high-salt diets (Cabral increased sorbitol and fructose production may be at higher risk
et al., 2014). Recent work suggests that fructose may also sensi- of developing T2D and its complications. Although sorbitol levels
tize the proximal tubule to angiotensin II and promote renal salt decrease in response to improved glycemic control in T2D, some
reabsorption (Gonzalez-Vicente et al., 2018; Yang et al., 2020). studies suggest that elevated sorbitol and fructose levels are
Additionally, KHK-mediated fructose metabolism within the more refractory to normalization of glycemia in some groups of
proximal tubule may enhance salt reabsorption by activating patients, again indicating that increased endogenous fructose
the sodium hydrogen exchanger (Slc9a3) (Hayasaki et al., production may contribute to the pathogenesis of T2D (Yoshii
2019; Queiroz-Leite et al., 2012). As fructolytic enzymes are ex- et al., 2001). Further studies are needed to better understand
pressed at high levels in the renal proximal tubule, the role of the causal relationship between circulating glucose concentra-
renal fructose metabolism as it relates to renal salt handling tions and endogenous fructose production. However, current
will likely be of growing interest. knowledge is compatible with a model whereby increased flux
through the polyol pathway, leading to increased endogenous
ENDOGENOUS PRODUCTION AS A MODIFIABLE fructose production, may present a significant, unappreciated
EXPOSURE? risk factor of developing T2D and its complications.
Studies in animal models also provide evidence of substantial
Although nearly all public health attention and most scientific endogenous fructose production. Indeed, micromolar circu-
research has focused on the metabolic effects of excessive die- lating fructose can be measured in GLUT5 knockout mice on
tary fructose, fructose is also synthesized endogenously from fructose-free diets (Patel et al., 2015c). Moreover, the inhibition
glucose through the sorbitol (polyol) pathway. In this pathway, of fructose metabolism by global deletion of KHK in mice leads
glucose is reduced by aldose reductase (AR, also known as to a 30-fold increase in circulating fructose on fructose-free diets
Akr1b1 in humans and Akr1b3 in mice) to sorbitol, which is (Patel et al., 2015c). Findings in animal models also support a
then oxidized to fructose by sorbitol dehydrogenase (SORD). pathologic role for endogenous fructose production in metabolic
Endogenously produced fructose appears to play a role in disease (Lanaspa et al., 2013, 2014; Oppelt et al., 2015). Mice
fertility, as it is a major energy source for sperm (Frenette et al., exposed to 10% glucose in drinking water develop features of
2006; Jayaraman et al., 2014; Martini et al., 2010). Sorbitol, the metabolic syndrome, including increased visceral adiposity, hy-
immediate precursor for fructose production, is also produced perinsulinemia, and fatty liver (Lanaspa et al., 2013). This effect is
by the placenta (Hwang et al., 2015). In fasted adult humans, blunted in both AR-deficient and KHK-deficient mice (Lanaspa
circulating plasma fructose is detectable in the low micromolar et al., 2013). As previously noted, mice with mutations in Aldob,
range, consistent with some degree of endogenous production, the enzyme required for catabolism of F1P, develop significant
and it tends to be higher in men than in women, and is possibly steatosis even on diets free of fructose, indicating that endoge-
related to production in the testes (Chen et al., 2020). nously produced fructose is sufficient to contribute to metabolic
Recently, Francey et al. measured endogenous fructose pro- dysfunction in this model (Oppelt et al., 2015).
duction rates of 5.6–16.7 g/day in vivo in healthy human partic- Sorbitol is a major intracellular osmolyte in the renal medulla,
ipants (Francey et al., 2019). Fructose production acutely where its production can be increased by hypertonicity through
increased with sugar consumption (Francey et al., 2019). An the induction of AR expression (Bagnasco et al., 1987). Theoret-
endogenous fructose production rate of 17 g/day is approxi- ically, high-salt diets might enhance endogenous fructose
mately one-third of the average daily consumption of fructose production within the kidney through this mechanism. Unexpect-
in the US population and thus a substantial fraction of total fruc- edly, Lanaspa and colleagues have observed that gavaging mice
tose exposure in many people (Marriott et al., 2009). Indeed, this with hypertonic fluids induced AR expression in the liver and hy-
is roughly equivalent to the fructose in a can of regular soda—an pothalamus and have proposed that endogenous fructose
amount that when consumed daily is associated with increased production is a conserved cellular program in response to hyper-
cardiometabolic risk (Green et al., 2014). tonicity in multiple tissues and that high-salt diets may contribute
Increased fasting fructosemia is associated with higher fasting to metabolic disease through this mechanism (Lanaspa et al.,
glucose and the risk of incident T2D (Chen et al., 2020). Because 2018b). Additional potential links between fructose metabolism,

2344 Cell Metabolism 33, December 7, 2021


ll
Review

the regulation of volume and hydration status, and metabolic dis- Abdelmalek, M.F., Suzuki, A., Guy, C., Unalp-Arida, A., Colvin, R., Johnson,
R.J., and Diehl, A.M.; Nonalcoholic Steatohepatitis Clinical Research Network
ease include the observation that fructose feeding can induce (2010). Increased fructose consumption is associated with fibrosis severity in
vasopressin secretion and that vasopressin activity through the patients with nonalcoholic fatty liver disease. Hepatology 51, 1961–1971.
vasopressin V1B receptor (Avpr1b) may contribute to the adverse
Abrigo, J.M., Shen, J., Wong, V.W.-S., Yeung, D.K.-W., Wong, G.L.-H., Chim,
metabolic effects of fructose (Andres-Hernando et al., 2021). A.M.-L., Chan, A.W.-H., Choi, P.C.-L., Chan, F.K.-L., Chan, H.L.-Y., and Chu-
W, W.C. (2014). Non-alcoholic fatty liver disease: spectral patterns observed
from an in vivo phosphorus magnetic resonance spectroscopy study.
CONCLUSIONS J. Hepatol 60, 809–815.

Aeberli, I., Hochuli, M., Gerber, P.A., Sze, L., Murer, S.B., Tappy, L., Spinas,
The last 5 years have witnessed major advancements in our un- G.A., and Berneis, K. (2013). Moderate amounts of fructose consumption
derstanding of the molecular physiology of fructose metabolism impair insulin sensitivity in healthy young men: a randomized controlled trial.
and its potential contribution to metabolic disease. One conclu- Diabetes Care 36, 150–156.

sion, firmly established in recent years, is that excessive fructose Agebratt, C., Ström, E., Romu, T., Dahlqvist-Leinhard, O., Borga, M., Lean-
metabolism within the liver is a major source of fructose-associ- dersson, P., and Nystrom, F.H. (2016). A randomized study of the effects of
additional fruit and nuts consumption on hepatic fat content, cardiovascular
ated morbidity. Over the next 5 years, we anticipate further prog- risk factors and basal metabolic rate. PLoS One 11, e0147149.
ress in defining key molecular mechanisms by which hepatic
fructose metabolism participates in disease pathogenesis. We Agius, L. (2008). Glucokinase and molecular aspects of liver glycogen meta-
bolism. Biochem. J. 414, 1–18.
are hopeful that mechanisms defined through these studies will
be useful not only to understand fructose-associated disease Aleman, J.O., Henderson, W.A., Walker, J.M., Ronning, A., Jones, D., Walter,
but may also generalize to other forms of metabolic disease asso- P.J., Daniel, S.G., Bittinger, K., Vaughan, R., MacArthur, R., et al. (2020).
Excess dietary fructose does not alter gut microbiota or permeability in hu-
ciated with obesity and overnutrition. Recent work has also es- mans: a randomized controlled pilot study. medRxiv. https://doi.org/10.
tablished that intestinal fructose metabolism may be protective. 1101/2020.11.23.20235515.
Defining mechanisms and pathways to enhance intestinal fruc- Alexander, C.M., Martin, J.A., Oxman, E., Kasza, I., Senn, K.A., and Dvinge, H.
tose metabolism seems a promising alternative or complemen- (2020). Alternative splicing and cleavage of GLUT8. Mol. Cell. Biol. 41.
tary therapeutic approach. Whereas fructose metabolism within
Ali, M., Rellos, P., and Cox, T.M. (1998). Hereditary fructose intolerance.
the intestine may be protective, the role of other fructose meta- J. Med. Genet. 35, 353–365.
bolism in other tissues, such as the kidney, remains to be firmly
Andres-Hernando, A., Kuwabara, M., Orlicky, D.J., Vandenbeuch, A., Cicerchi,
established. With the availability of conditional genetic mouse
C., Kinnamon, S.C., Finger, T.E., Johnson, R.J., and Lanaspa, M.A. (2020a).
models, we expect that this and related questions will be ad- Sugar causes obesity and metabolic syndrome in mice independently of sweet
dressed in short order. taste. Am. J. Physiol. Endocrinol. Metab. 319, E276–E290.
Advances in targeting fructose metabolism pharmacologically Andres-Hernando, A., Orlicky, D.J., Kuwabara, M., Ishimoto, T., Nakagawa, T.,
for therapeutic purposes have paralleled the advancements in Johnson, R.J., and Lanaspa, M.A. (2020b). Deletion of fructokinase in the liver
our understanding of its molecular physiology. Initial results in or in the intestine reveals differential effects on sugar-induced metabolic
dysfunction. Cell Metab 32, 117–127.e3.
early-stage clinical trials appear promising. Larger and longer-
term studies will be required to determine efficacy and safety Andres-Hernando, A., Jensen, T.J., Kuwabara, M., Orlicky, D.J., Cicerchi, C.,
Li, N., Roncal-Jimenez, C.A., Garcia, G.E., Ishimoto, T., Maclean, P.S., et al.
in diverse clinical settings with potential applications in popula-
(2021). Vasopressin mediates fructose-induced metabolic syndrome by acti-
tions with NAFLD and diabetes. vating the V1b receptor. JCI Insight 6, e140848.
The pharmacological and physiological data continue to sup-
Asipu, A., Hayward, B.E., O’Reilly, J., and Bonthron, D.T. (2003). Properties of
port recommendations to limit sugar consumption. Along with normal and mutant recombinant human ketohexokinases and implications for
the advancements in molecular physiology and pharmacology, the pathogenesis of essential fructosuria. Diabetes 52, 2426–2432.
establishing the efficacy of public health efforts to reduce sugar
Avena, N.M., Rada, P., and Hoebel, B.G. (2008). Evidence for sugar addiction:
consumption will remain an important goal for the foreseeable behavioral and neurochemical effects of intermittent, excessive sugar intake.
future. Neurosci. Biobehav. Rev. 32, 20–39.

Badve, S.V., Pascoe, E.M., Tiku, A., Boudville, N., Brown, F.G., Cass, A.,
ACKNOWLEDGMENTS Clarke, P., Dalbeth, N., Day, R.O., de Zoysa, J.R., et al. (2020). Effects of allo-
purinol on the progression of chronic kidney disease. N. Engl. J. Med 382,
2504–2513.
This work was supported by NIH grants R01DK100425 and 5R01DK121710
and the American Heart Association (16CSA28590003) (M.A.H.). Illustrations Baerlocher, K., Gitzelmann, R., Steinmann, B., and Gitzelmann-Cumarasamy,
created with BioRender.com. N. (1978). Hereditary fructose intolerance in early childhood: a major diag-
nostic challenge. Survey of 20 symptomatic cases. Helv. Paediatr. Acta 33,
465–487.
DECLARATION OF INTERESTS
Bagnasco, S.M., Uchida, S., Balaban, R.S., Kador, P.F., and Burg, M.B. (1987).
Induction of aldose reductase and sorbitol in renal inner medullary cells by
M.A.H. received research support from Eli Lilly and Co. M.J.B. is an employee
elevated extracellular NaCl. Proc. Natl. Acad. Sci. USA 84, 1718–1720.
and shareholder of Pfizer, which holds patent US20170183328A1.
Bailey, R.L., Fulgoni, V.L., Cowan, A.E., and Gaine, P.C. (2018). Sources of
added sugars in young children, adolescents, and adults with low and high in-
REFERENCES
takes of added sugars. Nutrients 10, 102.

Abdel Rahman, A., Jomaa, L., Kahale, L.A., Adair, P., and Pine, C. (2018). Barone, S., Fussell, S.L., Singh, A.K., Lucas, F., Xu, J., Kim, C., Wu, X., Yu, Y.,
Effectiveness of behavioral interventions to reduce the intake of sugar-sweet- Amlal, H., Seidler, U., et al. (2009). Slc2a5 (Glut5) is essential for the absorption
ened beverages in children and adolescents: a systematic review and meta- of fructose in the intestine and generation of fructose-induced hypertension.
analysis. Nutr. Rev. 76, 88–107. J. Biol. Chem. 284, 5056–5066.

Cell Metabolism 33, December 7, 2021 2345


ll
Review
Bawden, S.J., Stephenson, M.C., Ciampi, E., Hunter, K., Marciani, L., Mac- (2013). Novel locus including FGF21 is associated with dietary macronutrient
donald, I.A., Aithal, G.P., Morris, P.G., and Gowland, P.A. (2016). Investigating intake. Hum. Mol. Genet. 22, 1895–1902.
the effects of an oral fructose challenge on hepatic ATP reserves in healthy vol-
unteers: A (31)P MRS study. Clin. Nutr 35, 645–649. Chung, M., Ma, J., Patel, K., Berger, S., Lau, J., and Lichtenstein, A.H. (2014).
Fructose, high-fructose corn syrup, sucrose, and nonalcoholic fatty liver dis-
€mer, S., Volynets, V., Kaserouni, S.,
Bergheim, I., Weber, S., Vos, M., Kra ease or indexes of liver health: a systematic review and meta-analysis. Am.
McClain, C.J., and Bischoff, S.C. (2008). Antibiotics protect against fruc- J. Clin. Nutr 100, 833–849.
tose-induced hepatic lipid accumulation in mice: role of endotoxin.
J. Hepatol 48, 983–992. Coate, K.C., Scott, M., Farmer, B., Moore, M.C., Smith, M., Roop, J., Neal,
D.W., Williams, P., and Cherrington, A.D. (2010). Chronic consumption of a
Bismut, H., Hers, H.G., and Van Schaftingen, E. (1993). Conversion of fructose high-fat/high-fructose diet renders the liver incapable of net hepatic glucose
to glucose in the rabbit small intestine. A reappraisal of the direct pathway. Eur. uptake. Am. J. Physiol. Endocrinol. Metab. 299, E887–E898.
J. Biochem 213, 721–726.
Coate, K.C., Kraft, G., Moore, M.C., Smith, M.S., Ramnanan, C., Irimia, J.M.,
Bizeau, M.E., Thresher, J.S., and Pagliassotti, M.J. (2001). A high-sucrose diet Roach, P.J., Farmer, B., Neal, D.W., Williams, P., et al. (2014). Hepatic glucose
increases gluconeogenic capacity in isolated periportal and perivenous rat he- uptake and disposition during short-term high-fat vs. high-fructose feeding.
patocytes. Am. J. Physiol. Endocrinol. Metab. 280, E695–E702. Am. J. Physiol. Endocrinol. Metab. 307, E151–E160.

Bode, J.C., Zelder, O., Rumpelt, H.J., and Wittkamp, U. (1973). Depletion of Coffee and Caffeine Genetics Consortium, Cornelis, M.C., Byrne, E.M., Esko,
liver adenosine phosphates and metabolic effects of intravenous infusion of T., Nalls, M.A., Ganna, A., Paynter, N., Monda, K.L., Amin, N., Fischer, K., et al.
fructose or sorbitol in man and in the rat. Eur. J. Clin. Invest 3, 436–441. (2015). Genome-wide meta-analysis identifies six novel loci associated with
habitual coffee consumption. Mol. Psychiatry 20, 647–656.
Bray, G.A., Nielsen, S.J., and Popkin, B.M. (2004). Consumption of high-fruc-
tose corn syrup in beverages may play a role in the epidemic of obesity. Am. J. Concha, I.I., Velásquez, F.V., Martı́nez, J.M., Angulo, C., Droppelmann, A.,
Clin. Nutr 79, 537–543. Reyes, A.M., Slebe, J.C., Vera, J.C., and Golde, D.W. (1997). Human erythro-
cytes express GLUT5 and transport fructose. Blood 89, 4190–4195.
Bremer, A.A., Stanhope, K.L., Graham, J.L., Cummings, B.P., Wang, W., Sa-
ville, B.R., and Havel, P.J. (2011). Fructose-fed rhesus monkeys: a nonhuman Cook, G.C. (1969). Absorption products of D(-) fructose in man. Clin. Sci. 37,
primate model of insulin resistance, metabolic syndrome, and type 2 diabetes. 675–687.
Clin. Transl. Sci. 4, 243–252.
Cori, C.F. (1925). The fate of sugar in the animal body. J. Biol. Chem. 66,
Brownlee, M. (2005). The pathobiology of diabetic complications: a unifying 691–715.
mechanism. Diabetes 54, 1615–1625.
Cortez-Pinto, H., Chatham, J., Chacko, V.P., Arnold, C., Rashid, A., and Diehl,
Burant, C.F., Takeda, J., Brot-Laroche, E., Bell, G.I., and Davidson, N.O. A.M. (1999). Alterations in liver ATP homeostasis in human nonalcoholic stea-
(1992). Fructose transporter in human spermatozoa and small intestine is tohepatitis: a pilot study. JAMA 282, 1659–1664.
GLUT5. J. Biol. Chem. 267, 14523–14526.
Crossley, J.N., and Macdonald, I. (1970). The influence in male baboons, of a
Butler, A.A., Price, C.A., Graham, J.L., Stanhope, K.L., King, S., Hung, Y.H., high sucrose diet on the portal and arterial levels of glucose and fructose
Sethupathy, P., Wong, S., Hamilton, J., Krauss, R.M., et al. (2019). Fructose- following a sucrose meal. Nutr. Metab 12, 171–178.
induced hypertriglyceridemia in rhesus macaques is attenuated with fish oil
or ApoC3 RNA interference. J. Lipid Res. 60, 805–818. Curry, D.L. (1989). Effects of mannose and fructose on the synthesis and
secretion of insulin. Pancreas 4, 2–9.
Butler, A.A., Graham, J.L., Stanhope, K.L., Wong, S., King, S., Bremer, A.A.,
Krauss, R.M., Hamilton, J., and Havel, P.J. (2020). Role of angiopoietin-like Cydylo, M.A., Davis, A.T., and Kavanagh, K. (2017). Fatty liver promotes
protein 3 in sugar-induced dyslipidemia in rhesus macaques: suppression fibrosis in monkeys consuming high fructose. Obesity (Silver Spring) 25,
by fish oil or RNAi. J. Lipid Res. 61, 376–386. 290–293.

Cabral, P.D., Hong, N.J., Hye Khan, M.A., Ortiz, P.A., Beierwaltes, W.H., Imig, Davidson, N.O., Hausman, A.M., Ifkovits, C.A., Buse, J.B., Gould, G.W., Bur-
J.D., and Garvin, J.L. (2014). Fructose stimulates Na/H exchange activity and ant, C.F., and Bell, G.I. (1992). Human intestinal glucose transporter expres-
sensitizes the proximal tubule to angiotensin II. Hypertension 63, e68–e73. sion and localization of GLUT5. Am. J. Physiol 262, C795–C800.

Caron, S., Verrijken, A., Mertens, I., Samanez, C.H., Mautino, G., Haas, J.T., de Araujo, I.E., Oliveira-Maia, A.J., Sotnikova, T.D., Gainetdinov, R.R., Caron,
Duran-Sandoval, D., Prawitt, J., Francque, S., Vallez, E., et al. (2011). Tran- M.G., Nicolelis, M.A.L., and Simon, S.A. (2008). Food reward in the absence of
scriptional activation of apolipoprotein CIII expression by glucose may taste receptor signaling. Neuron 57, 930–941.
contribute to diabetic dyslipidemia. Arterioscler. Thromb. Vasc. Biol. 31,
513–519. de Koning, L., Malik, V.S., Kellogg, M.D., Rimm, E.B., Willett, W.C., and Hu,
F.B. (2012). Sweetened beverage consumption, incident coronary heart dis-
Chambers, R.A., and Pratt, R.T. (1956). Idiosyncrasy to fructose. Lancet ease, and biomarkers of risk in men. Circulation 125, 1735–1741, S1.
271, 340.
de Ruyter, J.C., Olthof, M.R., Seidell, J.C., and Katan, M.B. (2012). A trial of
Chambers, K.T., Chen, Z., Lai, L., Leone, T.C., Towle, H.C., Kralli, A., Crawford, sugar-free or sugar-sweetened beverages and body weight in children.
P.A., and Finck, B.N. (2013). PGC-1b and ChREBP partner to cooperatively N. Engl. J. Med 367, 1397–1406.
regulate hepatic lipogenesis in a glucose concentration-dependent manner.
Mol. Metab 2, 194–204. DeBosch, B.J., Chi, M., and Moley, K.H. (2012). Glucose transporter 8 (GLUT8)
regulates enterocyte fructose transport and global mammalian fructose utiliza-
Chen, Y., Lin, H., Qin, L., Lu, Y., Zhao, L., Xia, M., Jiang, J., Li, X., Yu, C., Zong, tion. Endocrinology 153, 4181–4191.
G., et al. (2020). Fasting serum fructose levels are associated with risk of inci-
dent type 2 diabetes in middle-aged and older Chinese population. Diabetes DeBosch, B.J., Chen, Z., Saben, J.L., Finck, B.N., and Moley, K.H. (2014).
Care 43, 2217–2225. Glucose transporter 8 (GLUT8) mediates fructose-induced de novo lipogen-
esis and macrosteatosis. J. Biol. Chem. 289, 10989–10998.
Chiu, S., Sievenpiper, J.L., de Souza, R.J., Cozma, A.I., Mirrahimi, A., Carleton,
A.J., Ha, V., Di Buono, M., Jenkins, A.L., Leiter, L.A., et al. (2014). Effect of fruc- Debray, F.G., Seyssel, K., Fadeur, M., Tappy, L., Paquot, N., and Tran, C.
tose on markers of non-alcoholic fatty liver disease (NAFLD): a systematic re- (2021). Effect of a high fructose diet on metabolic parameters in carriers for he-
view and meta-analysis of controlled feeding trials. Eur. J. Clin. Nutr 68, reditary fructose intolerance. Clin. Nutr 40, 4246–4254.
416–423.
Dencker, H., Lunderquist, A., Meeuwisse, G., Norryd, C., and Tranberg, K.G.
Chotiwat, C., Sharp, C., Teff, K., and Harris, R.B.S. (2007). Feeding a high-fruc- (1972). Absorption of fructose as measured by portal catheterization. Scand.
tose diet induces leptin resistance in rats. Appetite 49, 284. J. Gastroenterol 7, 701–705.

Chu, A.Y., Workalemahu, T., Paynter, N.P., Rose, L.M., Giulianini, F., Tanaka, Dentin, R., Pégorier, J.P., Benhamed, F., Foufelle, F., Ferré, P., Fauveau, V.,
T., Ngwa, J.S., CHARGE Nutrition Working Group, Qi, Q., Curhan, G.C., et al. Magnuson, M.A., Girard, J., and Postic, C. (2004). Hepatic glucokinase is

2346 Cell Metabolism 33, December 7, 2021


ll
Review
required for the synergistic action of ChREBP and SREBP-1c on glycolytic and Exton, J.H., Jefferson, L.S., Butcher, R.W., and Park, C.R. (1966). Gluconeo-
lipogenic gene expression. J. Biol. Chem. 279, 20314–20326. genesis in the perfused liver. The effects of fasting, alloxan diabetes, glucagon,
epinephrine, adenosine 30 ,50 -monophosphate and insulin. Am. J. Med 40,
Dhingra, R., Sullivan, L., Jacques, P.F., Wang, T.J., Fox, C.S., Meigs, J.B., 709–715.
D’Agostino, R.B., Gaziano, J.M., and Vasan, R.S. (2007). Soft drink consump-
tion and risk of developing cardiometabolic risk factors and the metabolic syn- Farquhar, J.W., Frank, A., Gross, R.C., and Reaven, G.M. (1966). Glucose, in-
drome in middle-aged adults in the community. Circulation 116, 480–488. sulin, and triglyceride responses to high and low carbohydrate diets in man.
J. Clin. Invest 45, 1648–1656.
Diggle, C.P., Shires, M., Leitch, D., Brooke, D., Carr, I.M., Markham, A.F., Hay-
ward, B.E., Asipu, A., and Bonthron, D.T. (2009). Ketohexokinase: expression Fernandez, M.A., and Raine, K.D. (2019). Insights on the influence of sugar
and localization of the principal fructose-metabolizing enzyme. J. Histochem. taxes on obesity prevention efforts. Curr. Nutr. Rep 8, 333–339.
Cytochem 57, 763–774.
Fernández-Bañares, F., Esteve-Pardo, M., de Leon, R., Humbert, P., Cabré,
Diggle, C.P., Shires, M., McRae, C., Crellin, D., Fisher, J., Carr, I.M., Markham, E., Llovet, J.M., and Gassull, M.A. (1993). Sugar malabsorption in functional
A.F., Hayward, B.E., Asipu, A., Bonthron, D.T., et al. (2010). Both isoforms of bowel disease: clinical implications. Am. J. Gastroenterol 88, 2044–2050.
ketohexokinase are dispensable for normal growth and development. Physiol.
Genomics 42A, 235–243. Ferraris, R.P. (2001). Dietary and developmental regulation of intestinal sugar
transport. Biochem. J. 360, 265–276.
Donnelly, K.L., Smith, C.I., Schwarzenberg, S.J., Jessurun, J., Boldt, M.D., and
Parks, E.J. (2005). Sources of fatty acids stored in liver and secreted via lipo- Ferraris, R.P., Choe, J.Y., and Patel, C.R. (2018). Intestinal absorption of fruc-
proteins in patients with nonalcoholic fatty liver disease. J. Clin. Invest 115, tose. Annu. Rev. Nutr 38, 41–67.
1343–1351.
Flamment, M., Hajduch, E., Ferré, P., and Foufelle, F. (2012). New insights into
Doria, A., Galecki, A.T., Spino, C., Pop-Busui, R., Cherney, D.Z., Lingvay, I., ER stress-induced insulin resistance. Trends Endocrinol. Metab 23, 381–390.
Parsa, A., Rossing, P., Sigal, R.J., Afkarian, M., et al. (2020). Serum urate
lowering with allopurinol and kidney function in type 1 diabetes. N. Engl. J. Flanagan, E.W., Beyl, R.A., Fearnbach, S.N., Altazan, A.D., Martin, C.K., and
Med 382, 2493–2503. Redman, L.M. (2021). The impact of COVID-19 stay-at-home orders on health
behaviors in adults. Obesity (Silver Spring) 29, 438–445.
Dotimas, J.R., Lee, A.W., Schmider, A.B., Carroll, S.H., Shah, A., Bilen, J., El-
liott, K.R., Myers, R.B., Soberman, R.J., Yoshioka, J., and Lee, R.T. (2016). Flippo, K.H., and Potthoff, M.J. (2021). Metabolic messengers: FGF21. Nat.
Diabetes regulates fructose absorption through thioredoxin-interacting pro- Metab 3, 309–317.
tein. eLife 5, e18313.
Francey, C., Cros, J., Rosset, R., Crézé, C., Rey, V., Stefanoni, N., Schneiter,
Douard, V., and Ferraris, R.P. (2008). Regulation of the fructose transporter P., Tappy, L., and Seyssel, K. (2019). The extra-splanchnic fructose escape af-
GLUT5 in health and disease. Am. J. Physiol. Endocrinol. Metab. 295, ter ingestion of a fructose-glucose drink: an exploratory study in healthy hu-
E227–E237. mans using a dual fructose isotope method. Clin. Nutr. ESPEN 29, 125–132.
Dushay, J.R., Toschi, E., Mitten, E.K., Fisher, F.M., Herman, M.A., and Mara- Frenette, G., Thabet, M., and Sullivan, R. (2006). Polyol pathway in human
tos-Flier, E. (2015). Fructose ingestion acutely stimulates circulating FGF21 epididymis and semen. J. Androl 27, 233–239.
levels in humans. Mol. Metab 4, 51–57.
Froesch, E.R., and Ginsberg, J.L. (1962). Fructose metabolism of adipose tis-
Dyer, J., Salmon, K.S.H., Zibrik, L., and Shirazi-Beechey, S.P. (2005). Expres-
sue. I. Comparison of fructose and glucose metabolism in epididymal adipose
sion of sweet taste receptors of the T1R family in the intestinal tract and enter-
tissue of normal rats. J. Biol. Chem. 237, 3317–3324.
oendocrine cells. Biochem. Soc. Trans. 33, 302–305.
Fu, Z., Berhane, F., Fite, A., Seyoum, B., Abou-Samra, A.B., and Zhang, R.
Eaton, R.P., and Kipnis, D.M. (1969). Effects of high-carbohydrate diets on
(2014). Elevated circulating lipasin/betatrophin in human type 2 diabetes and
lipid and carbohydrate metabolism in the rat. Am. J. Physiol 217, 1160–1168.
obesity. Sci. Rep 4, 5013.
Ebbeling, C.B., Feldman, H.A., Chomitz, V.R., Antonelli, T.A., Gortmaker, S.L.,
Osganian, S.K., and Ludwig, D.S. (2012). A randomized trial of sugar-sweet- Futatsugi, K., Smith, A.C., Tu, M., Raymer, B., Ahn, K., Coffey, S.B., Dowling,
ened beverages and adolescent body weight. N. Engl. J. Med 367, 1407–1416. M.S., Fernando, D.P., Gutierrez, J.A., Huard, K., et al. (2020). Discovery of PF-
06835919: a potent inhibitor of ketohexokinase (KHK) for the treatment of
Ebbeling, C.B., Feldman, H.A., Steltz, S.K., Quinn, N.L., Robinson, L.M., and metabolic disorders driven by the overconsumption of fructose. J. Med.
Ludwig, D.S. (2020). Effects of sugar-sweetened, artificially sweetened, and Chem. 63, 13546–13560.
unsweetened beverages on cardiometabolic risk factors, body composition,
and sweet taste preference: a randomized controlled trial. J. Am. Heart Assoc. Ge, J.Y., Ji, Y., Zhu, Z.Y., and Li, X. (2020). Genetically elevated serum uric acid
9, e015668. and renal function in an apparently healthy population. Urol. Int 104, 277–282.

Eggleston, L.V., and Woods, H.F. (1970). Activation of liver pyruvate kinase by Geidl-Flueck, B., Hochuli, M., Németh, Á., Eberl, A., Derron, N., Köfeler, H.C.,
fructose-1-phosphate. FEBS Lett. 6, 43–45. Tappy, L., Berneis, K., Spinas, G.A., and Gerber, P.A. (2021). Fructose- and su-
crose- but not glucose-sweetened beverages promote hepatic de novo lipo-
Eissing, L., Scherer, T., Tödter, K., Knippschild, U., Greve, J.W., Buurman, genesis: a randomized controlled trial. J. Hepatol 75, 46–54.
W.A., Pinnschmidt, H.O., Rensen, S.S., Wolf, A.M., Bartelt, A., et al. (2013).
De novo lipogenesis in human fat and liver is linked to ChREBP-b and meta- Gibson, P.R., and Shepherd, S.J. (2005). Personal view: food for thought–
bolic health. Nat. Commun 4, 1528. western lifestyle and susceptibility to Crohn’s disease. The FODMAP hypoth-
esis. Aliment. Pharmacol. Ther 21, 1399–1409.
Ejaz, A.A., Nakagawa, T., Kanbay, M., Kuwabara, M., Kumar, A., Garcia
Arroyo, F.E., Roncal-Jimenez, C., Sasai, F., Kang, D.H., Jensen, T., et al. Ginsburg, V., and Hers, H.G. (1960). On the conversion of fructose to glucose
(2020). Hyperuricemia in kidney disease: a major risk factor for cardiovascular by guinea pig intestine. Biochim. Biophys. Acta 38, 427–434.
events, vascular calcification, and renal damage. Semin. Nephrol 40, 574–585.
Gonzalez-Vicente, A., Hong, N.J., Yang, N., Cabral, P.D., Berthiaume, J.M.,
Emerson, H., and Larimore, L.D. (1924). Diabetes mellitus: a contribution to its Dominici, F.P., and Garvin, J.L. (2018). Dietary fructose increases the sensi-
epidemiology based chiefly on mortality statistics. Arch. Intern. Med. (Chic) 34, tivity of proximal tubules to angiotensin II in rats fed high-salt diets. Nutri-
585–630. ents 10.

Erion, D.M., Popov, V., Hsiao, J.J., Vatner, D., Mitchell, K., Yonemitsu, S., Gopher, A., Vaisman, N., Mandel, H., and Lapidot, A. (1990). Determination of
Nagai, Y., Kahn, M., Gillum, M.P., Dong, J., et al. (2013). The role of the carbo- fructose metabolic pathways in normal and fructose-intolerant children: a 13C
hydrate response element-binding protein in male fructose-fed rats. Endocri- NMR study using fructose. Proc. Natl. Acad. Sci. USA 87, 5449–5453.
nology 154, 36–44.
Gorboulev, V., Schu €rmann, A., Vallon, V., Kipp, H., Jaschke, A., Klessen, D.,
Exton, J.H., and Park, C.R. (1967). Control of gluconeogenesis in liver. I. Gen- Friedrich, A., Scherneck, S., Rieg, T., Cunard, R., et al. (2012). Na(+)-D-glucose
eral features of gluconeogenesis in the perfused livers of rats. J. Biol. Chem. cotransporter SGLT1 is pivotal for intestinal glucose absorption and glucose-
242, 2622–2636. dependent incretin secretion. Diabetes 61, 187–196.

Cell Metabolism 33, December 7, 2021 2347


ll
Review
Green, A.K., Jacques, P.F., Rogers, G., Fox, C.S., Meigs, J.B., and McKeown, Hudgins, L.C., Parker, T.S., Levine, D.M., and Hellerstein, M.K. (2011). A dual
N.M. (2014). Sugar-sweetened beverages and prevalence of the metabolically sugar challenge test for lipogenic sensitivity to dietary fructose. J. Clin. Endo-
abnormal phenotype in the Framingham Heart Study. Obesity (Silver Spring) crinol. Metab 96, 861–868.
22, E157–E163.
Hwang, I.S., Ho, H., Hoffman, B.B., and Reaven, G.M. (1987). Fructose-
Grossbard, L., and Schimke, R.T. (1966). Multiple hexokinases of rat tissues. induced insulin resistance and hypertension in rats. Hypertension 10, 512–516.
Purification and comparison of soluble forms. J. Biol. Chem. 241, 3546–3560.
Hwang, J.J., Johnson, A., Cline, G., Belfort-Deaguiar, R., Snegovskikh, D.,
Gutierrez, J.A., Liu, W., Perez, S., Xing, G., Sonnenberg, G., Kou, K., Blatnik, Khokhar, B., Han, C.S., and Sherwin, R.S. (2015). Fructose levels are markedly
M., Allen, R., Weng, Y., Vera, N.B., et al. (2021). Pharmacologic inhibition of ke- elevated in cerebrospinal fluid compared with plasma in pregnant women.
tohexokinase prevents fructose-induced metabolic dysfunction. Mol. Metab PLoS One 10, e0128582.
48, 101196.
Iizuka, K., Bruick, R.K., Liang, G., Horton, J.D., and Uyeda, K. (2004). Defi-
ciency of carbohydrate response element-binding protein (ChREBP) reduces
Haas, J.T., Miao, J., Chanda, D., Wang, Y., Zhao, E., Haas, M.E., Hirschey, M.,
lipogenesis as well as glycolysis. Proc. Natl. Acad. Sci. USA 101, 7281–7286.
Vaitheesvaran, B., Farese, R.V., Kurland, I.J., et al. (2012). Hepatic insulin
signaling is required for obesity-dependent expression of SREBP-1c mRNA Iizuka, K., Takeda, J., and Horikawa, Y. (2009). Glucose induces FGF21 mRNA
but not for feeding-dependent expression. Cell Metab 15, 873–884. expression through ChREBP activation in rat hepatocytes. FEBS Lett. 583,
2882–2886.
Haidari, M., Leung, N., Mahbub, F., Uffelman, K.D., Kohen-Avramoglu, R.,
Lewis, G.F., and Adeli, K. (2002). Fasting and postprandial overproduction of Imamura, F., O’Connor, L., Ye, Z., Mursu, J., Hayashino, Y., Bhupathiraju, S.N.,
intestinally derived lipoproteins in an animal model of insulin resistance. Evi- and Forouhi, N.G. (2015). Consumption of sugar sweetened beverages, artifi-
dence that chronic fructose feeding in the hamster is accompanied by cially sweetened beverages, and fruit juice and incidence of type 2 diabetes:
enhanced intestinal de novo lipogenesis and ApoB48-containing lipoprotein systematic review, meta-analysis, and estimation of population attributable
overproduction. J. Biol. Chem. 277, 31646–31655. fraction. BMJ 351, h3576.

Han, W., Tellez, L.A., Perkins, M.H., Perez, I.O., Qu, T., Ferreira, J., Ferreira, Ishimoto, T., Lanaspa, M.A., Le, M.T., Garcia, G.E., Diggle, C.P., Maclean,
T.L., Quinn, D., Liu, Z.-W., Gao, X.-B., et al. (2018). A neural circuit for gut- P.S., Jackman, M.R., Asipu, A., Roncal-Jimenez, C.A., Kosugi, T., et al.
induced reward. Cell 175, 665–678.e23. (2012). Opposing effects of fructokinase C and A isoforms on fructose-induced
metabolic syndrome in mice. Proc. Natl. Acad. Sci. USA 109, 4320–4325.
Hasan, N.M., Johnson, K.F., Yin, J., Baetz, N.W., Fayad, L., Sherman, V., Blutt,
S.E., Estes, M.K., Kumbhari, V., Zachos, N.C., and Kovbasnjuk, O. (2021). In- Jamnik, J., Rehman, S., Blanco Mejia, S., de Souza, R.J., Khan, T.A., Leiter,
testinal stem cell-derived enteroids from morbidly obese patients preserve L.A., Wolever, T.M.S., Kendall, C.W.C., Jenkins, D.J.A., and Sievenpiper,
obesity-related phenotypes: elevated glucose absorption and gluconeogen- J.L. (2016). Fructose intake and risk of gout and hyperuricemia: a systematic
esis. Mol. Metab 44, 101129. review and meta-analysis of prospective cohort studies. BMJ Open 6,
e013191.
Haslam, D.E., Peloso, G.M., Herman, M.A., Dupuis, J., Lichtenstein, A.H.,
Smith, C.E., and McKeown, N.M. (2020). Beverage consumption and longitu- Jang, C., Hui, S., Lu, W., Cowan, A.J., Morscher, R.J., Lee, G., Liu, W., Tesz,
dinal changes in lipoprotein concentrations and incident dyslipidemia in US G.J., Birnbaum, M.J., and Rabinowitz, J.D. (2018). The small intestine converts
adults: the Framingham heart study. J. Am. Heart Assoc. 9, e014083. dietary fructose into glucose and organic acids. Cell Metab 27, 351–361.e3.

Jang, C., Wada, S., Yang, S., Gosis, B., Zeng, X., Zhang, Z., Shen, Y., Lee, G.,
Haslam, D.E., Peloso, G.M., Guirette, M., Imamura, F., Bartz, T.M., Pitsillides,
Arany, Z., and Rabinowitz, J.D. (2020). The small intestine shields the liver from
€nen, N., et al. (2021). Sugar-
A.N., Wang, C.A., Li-Gao, R., Westra, J.M., Pitka
fructose-induced steatosis. Nat. Metab 2, 586–593.
sweetened beverage consumption may modify associations between genetic
variants in the CHREBP (carbohydrate responsive element binding protein) lo- Jayalath, V.H., de Souza, R.J., Ha, V., Mirrahimi, A., Blanco-Mejia, S., Di
cus and HDL-C (high-density lipoprotein cholesterol) and triglyceride concen- Buono, M., Jenkins, A.L., Leiter, L.A., Wolever, T.M., Beyene, J., et al.
trations. Circ. Genom. Precis. Med. 14, e003288. (2015). Sugar-sweetened beverage consumption and incident hypertension:
a systematic review and meta-analysis of prospective cohorts. Am. J. Clin.
Hauri, H.P., Quaroni, A., and Isselbacher, K.J. (1979). Biogenesis of intestinal Nutr 102, 914–921.
plasma membrane: posttranslational route and cleavage of sucrase-isomal-
tase. Proc. Natl. Acad. Sci. USA 76, 5183–5186. Jayaraman, V., Ghosh, S., SenGupta, A., Srivastava, S., Sonawat, H.M., and
Narayan, P.K. (2014). Identification of biochemical differences between
Hayasaki, T., Ishimoto, T., Doke, T., Hirayama, A., Soga, T., Furuhashi, K., different forms of male infertility by nuclear magnetic resonance (NMR) spec-
Kato, N., Kosugi, T., Tsuboi, N., Lanaspa, M.A., et al. (2019). Fructose in- troscopy. J. Assist. Reprod. Genet. 31, 1195–1204.
creases the activity of sodium hydrogen exchanger in renal proximal tubules
that is dependent on ketohexokinase. J. Nutr. Biochem 71, 54–62. Jensen, T., Abdelmalek, M.F., Sullivan, S., Nadeau, K.J., Green, M., Roncal,
C., Nakagawa, T., Kuwabara, M., Sato, Y., Kang, D.-H., et al. (2018). Fructose
Hayward, B.E., and Bonthron, D.T. (1998). Structure and alternative splicing of and sugar: a major mediator of non-alcoholic fatty liver disease. J. Hepatol 68,
the ketohexokinase gene. Eur. J. Biochem 257, 85–91. 1063–1075.

Heinz, F., Lamprecht, W., and Kirsch, J. (1968). Enzymes of fructose meta- Jensen-Cody, S.O., Flippo, K.H., Claflin, K.E., Yavuz, Y., Sapouckey, S.A.,
bolism in human liver. J. Clin. Invest 47, 1826–1832. Walters, G.C., Usachev, Y.M., Atasoy, D., Gillum, M.P., and Potthoff, M.J.
(2020). FGF21 signals to glutamatergic neurons in the ventromedial hypothal-
Hers, H.G. (1957). Le Métabolisme du Fructose (UCL - Université Catholique amus to suppress carbohydrate intake. Cell Metab 32, 273–286.e6.
de Louvain).
Johnson, R.K., Appel, L.J., Brands, M., Howard, B.V., Lefevre, M., Lustig, R.H.,
Sacks, F., Steffen, L.M., and Wylie-Rosett, J.; American Heart Association
Hers, H.G., and Joassin, G. (1961). Anomaly of hepatic aldolase in intolerance
Nutrition Committee of the Council on Nutrition, Physical Activity, and Meta-
to fructose. Enzymol. Biol. Clin. (Basel) 1, 4–14.
bolism and the Council on Epidemiology and Prevention (2009). Dietary sugars
intake and cardiovascular health: a scientific statement from the American
Hieronimus, B., Medici, V., Bremer, A.A., Lee, V., Nunez, M.V., Sigala, D.M.,
Heart Association. Circulation 120, 1011–1020.
Keim, N.L., Havel, P.J., and Stanhope, K.L. (2020). Synergistic effects of fruc-
tose and glucose on lipoprotein risk factors for cardiovascular disease in Johnson, R.J., Sánchez-Lozada, L.G., Andrews, P., and Lanaspa, M.A. (2017).
young adults. Metabolism 112, 154356. Perspective: a historical and scientific perspective of sugar and its relation with
obesity and diabetes. Adv. Nutr 8, 412–422.
Holdsworth, C.D., and Dawson, A.M. (1965). Absorption of fructose in man.
Proc. Soc. Exp. Biol. Med 118, 142–145. Johnston, J.A., Nelson, D.R., Bhatnagar, P., Curtis, S.E., Chen, Y., and Mack-
rell, J.G. (2021). Prevalence and cardiometabolic correlates of ketohexokinase
Hu, Y., Semova, I., Sun, X., Kang, H., Chahar, S., Hollenberg, A.N., Masson, D., gene variants among UK Biobank participants. PLoS One 16, e0247683.
Hirschey, M.D., Miao, J., and Biddinger, S.B. (2018). Fructose and glucose can
regulate mammalian target of rapamycin complex 1 and lipogenic gene Jordan, D.M., Choi, H.K., Verbanck, M., Topless, R., Won, H.H., Nadkarni, G.,
expression via distinct pathways. J. Biol. Chem. 293, 2006–2014. Merriman, T.R., and Do, R. (2019). No causal effects of serum urate levels on

2348 Cell Metabolism 33, December 7, 2021


ll
Review
the risk of chronic kidney disease: a Mendelian randomization study. PLoS Kooner, J.S., Chambers, J.C., Aguilar-Salinas, C.A., Hinds, D.A., Hyde, C.L.,
Med 16, e1002725. Warnes, G.R., Gómez Pérez, F.J., Frazer, K.A., Elliott, P., Scott, J., et al.
(2008). Genome-wide scan identifies variation in MLXIPL associated with
Jurczak, M.J., Lee, A.H., Jornayvaz, F.R., Lee, H.Y., Birkenfeld, A.L., Guigni, plasma triglycerides. Nat. Genet. 40, 149–151.
B.A., Kahn, M., Samuel, V.T., Glimcher, L.H., and Shulman, G.I. (2012). Disso-
ciation of inositol-requiring enzyme (IRE1a)-mediated c-Jun N-terminal kinase Kursawe, R., Caprio, S., Giannini, C., Narayan, D., Lin, A., D’Adamo, E., Shaw,
activation from hepatic insulin resistance in conditional X-box-binding protein- M., Pierpont, B., Cushman, S.W., and Shulman, G.I. (2013). Decreased tran-
1 (XBP1) knock-out mice. J. Biol. Chem. 287, 2558–2567. scription of ChREBP-a/b isoforms in abdominal subcutaneous adipose tissue
of obese adolescents with prediabetes or early type 2 diabetes: associations
Kaelberer, M.M., Buchanan, K.L., Klein, M.E., Barth, B.B., Montoya, M.M., with insulin resistance and hyperglycemia. Diabetes 62, 837–844.
Shen, X., and Bohórquez, D.V. (2018). A gut-brain neural circuit for nutrient
sensory transduction. Science 361, eaat5236. Lambert, J.E., Ramos-Roman, M.A., Browning, J.D., and Parks, E.J. (2014).
Increased de novo lipogenesis is a distinct characteristic of individuals with
Kammoun, H.L., Chabanon, H., Hainault, I., Luquet, S., Magnan, C., Koike, T., nonalcoholic fatty liver disease. Gastroenterology 146, 726–735.
Ferré, P., and Foufelle, F. (2009). GRP78 expression inhibits insulin and ER
stress-induced SREBP-1c activation and reduces hepatic steatosis in mice. Lambertz, J., Weiskirchen, S., Landert, S., and Weiskirchen, R. (2017). Fruc-
J. Clin. Invest 119, 1201–1215. tose: a dietary sugar in crosstalk with microbiota contributing to the develop-
ment and progression of non-alcoholic liver disease. Front. Immunol 8, 1159.
Kanerva, N., Sandboge, S., Kaartinen, N.E., Ma €nnistö, S., and Eriksson, J.G.
Lanaspa, M.A., Sanchez-Lozada, L.G., Choi, Y.J., Cicerchi, C., Kanbay, M.,
(2014). Higher fructose intake is inversely associated with risk of nonalcoholic
Roncal-Jimenez, C.A., Ishimoto, T., Li, N., Marek, G., Duranay, M., et al.
fatty liver disease in older Finnish adults. Am. J. Clin. Nutr 100, 1133–1138.
(2012). Uric acid induces hepatic steatosis by generation of mitochondrial
oxidative stress: potential role in fructose-dependent and -independent fatty
Karczmar, G.S., Kurtz, T., Tavares, N.J., and Weiner, M.W. (1989). Regulation
liver. J. Biol. Chem. 287, 40732–40744.
of hepatic inorganic phosphate and ATP in response to fructose loading: an
in vivo 31P-NMR study. Biochim. Biophys. Acta 1012, 121–127. 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. (2013). Endog-
Kato, T., Iizuka, K., Takao, K., Horikawa, Y., Kitamura, T., and Takeda, J. enous fructose production and metabolism in the liver contributes to the devel-
(2018). ChREBP-knockout mice show sucrose intolerance and fructose opment of metabolic syndrome. Nat. Commun 4, 2434.
malabsorption. Nutrients 10.
Lanaspa, M.A., Ishimoto, T., Cicerchi, C., Tamura, Y., Roncal-Jimenez, C.A.,
Katz, L.S., Baumel-Alterzon, S., Scott, D.K., and Herman, M.A. (2021). Adap- Chen, W., Tanabe, K., Andres-Hernando, A., Orlicky, D.J., Finol, E., et al.
tive and maladaptive roles for ChREBP in the liver and pancreatic islets. J. Biol. (2014). Endogenous fructose production and fructokinase activation mediate
Chem. 296, 100623. renal injury in diabetic nephropathy. J. Am. Soc. Nephrol 25, 2526–2538.
Kaufmann, U., and Froesch, E.R. (1973). Inhibition of phosphorylase-a by fruc- Lanaspa, M.A., Andres-Hernando, A., Orlicky, D.J., Cicerchi, C., Jang, C., Li,
tose-1-phosphate, alpha-glycerophosphate and fructose-1,6-diphosphate: N., Milagres, T., Kuwabara, M., Wempe, M.F., Rabinowitz, J.D., et al.
explanation for fructose-induced hypoglycaemia in hereditary fructose intoler- (2018a). Ketohexokinase C blockade ameliorates fructose-induced metabolic
ance and fructose-1,6-diphosphatase deficiency. Eur. J. Clin. Invest 3, dysfunction in fructose-sensitive mice. J. Clin. Invest 128, 2226–2238.
407–413.
Lanaspa, M.A., Kuwabara, M., Andres-Hernando, A., Li, N., Cicerchi, C., Jen-
Kavanagh, K., Wylie, A.T., Tucker, K.L., Hamp, T.J., Gharaibeh, R.Z., Fodor, sen, T., Orlicky, D.J., Roncal-Jimenez, C.A., Ishimoto, T., Nakagawa, T., et al.
A.A., and Cullen, J.M.C. (2013). Dietary fructose induces endotoxemia and he- (2018b). High salt intake causes leptin resistance and obesity in mice by stim-
patic injury in calorically controlled primates. Am. J. Clin. Nutr 98, 349–357. ulating endogenous fructose production and metabolism. Proc. Natl. Acad.
Sci. U. S. A 115, 3138–3143.
Kayano, T., Burant, C.F., Fukumoto, H., Gould, G.W., Fan, Y.S., Eddy, R.L.,
Byers, M.G., Shows, T.B., Seino, S., and Bell, G.I. (1990). Human facilitative Landau, B.R., Marshall, J.S., Craig, J.W., Hostetler, K.Y., and Genuth, S.M.
glucose transporters. Isolation, functional characterization, and gene localiza- (1971). Quantitation of the pathways of fructose metabolism in normal and
tion of cDNAs encoding an isoform (GLUT5) expressed in small intestine, fructose-intolerant subjects. J. Lab. Clin. Med 78, 608–618.
kidney, muscle, and adipose tissue and an unusual glucose transporter pseu-
dogene-like sequence (GLUT6). J. Biol. Chem. 265, 13276–13282. Laron, Z. (1961). Essential benign fructosuria. Arch. Dis. Child 36, 273–277.

Kazierad, D.J., Chidsey, K., Somayaji, V.R., Bergman, A.J., Birnbaum, M.J., Lê, K.A., Ith, M., Kreis, R., Faeh, D., Bortolotti, M., Tran, C., Boesch, C., and
and Calle, R.A. (2021). Inhibition of ketohexokinase in adults with NAFLD re- Tappy, L. (2009). Fructose overconsumption causes dyslipidemia and ectopic
duces liver fat and inflammatory markers: a randomized phase 2 trial. Med lipid deposition in healthy subjects with and without a family history of type 2
2, 800–813.e3. diabetes. Am. J. Clin. Nutr 89, 1760–1765.

Khan, T.A., and Sievenpiper, J.L. (2016). Controversies about sugars: results Le, M.T., Frye, R.F., Rivard, C.J., Cheng, J., McFann, K.K., Segal, M.S., John-
from systematic reviews and meta-analyses on obesity, cardiometabolic dis- son, R.J., and Johnson, J.A. (2012). Effects of high-fructose corn syrup and
ease and diabetes. Eur. J. Nutr 55, 25–43. sucrose on the pharmacokinetics of fructose and acute metabolic and hemo-
dynamic responses in healthy subjects. Metabolism 61, 641–651.
Kim, H.R., Park, S.W., Cho, H.J., Chae, K.A., Sung, J.M., Kim, J.S., Landowski,
Le, M.T., Lanaspa, M.A., Cicerchi, C.M., Rana, J., Scholten, J.D., Hunter, B.L.,
C.P., Sun, D., Abd El-Aty, A.M., Amidon, G.L., and Shin, H.-C. (2007). Compar-
Rivard, C.J., Randolph, R.K., and Johnson, R.J. (2016). Bioactivity-guided
ative gene expression profiles of intestinal transporters in mice, rats and hu-
identification of botanical inhibitors of ketohexokinase. PLoS One 11,
mans. Pharmacol. Res. 56, 224–236.
e0157458.
Kim, M.S., Krawczyk, S.A., Doridot, L., Fowler, A.J., Wang, J.X., Trauger, S.A., Lee, A.H., Scapa, E.F., Cohen, D.E., and Glimcher, L.H. (2008). Regulation of
Noh, H.L., Kang, H.J., Meissen, J.K., Blatnik, M., et al. (2016). ChREBP regu- hepatic lipogenesis by the transcription factor XBP1. Science 320, 1492–1496.
lates fructose-induced glucose production independently of insulin signaling.
J. Clin. Invest 126, 4372–4386. Lei, Y., Hoogerland, J.A., Bloks, V.W., Bos, T., Bleeker, A., Wolters, H., Wol-
ters, J.C., Hijmans, B.S., van Dijk, T.H., Thomas, R., et al. (2020). Hepatic car-
Kim, M., Astapova, I.I., Flier, S.N., Hannou, S.A., Doridot, L., Sargsyan, A., Kou, bohydrate response element binding protein activation limits nonalcoholic
H.H., Fowler, A.J., Liang, G., and Herman, M.A. (2017). Intestinal, but not he- fatty liver disease development in a mouse model for glycogen storage disease
patic, ChREBP is required for fructose tolerance. JCI Insight 2, e96703. type 1a. Hepatology 72, 1638–1653.

Kim, J.Y., Garcia-Carbonell, R., Yamachika, S., Zhao, P., Dhar, D., Loomba, Leigh, S.J., and Morris, M.J. (2018). The role of reward circuitry and food
R., Kaufman, R.J., Saltiel, A.R., and Karin, M. (2018). ER stress drives lipogen- addiction in the obesity epidemic: an update. Biol. Psychol 131, 31–42.
esis and steatohepatitis via caspase-2 activation of S1P. Cell 175, 133–
145.e15. Li, X., Qian, X., Peng, L.-X., Jiang, Y., Hawke, D.H., Zheng, Y., Xia, Y., Lee, J.-
H., Cote, G., Wang, H., et al. (2016). A splicing switch from ketohexokinase-C
Klein, A.V., and Kiat, H. (2015). The mechanisms underlying fructose-induced to ketohexokinase-A drives hepatocellular carcinoma formation. Nat. Cell Biol.
hypertension: a review. J. Hypertens 33, 912–920. 18, 561–571.

Cell Metabolism 33, December 7, 2021 2349


ll
Review
Lim, J.S., Mietus-Snyder, M., Valente, A., Schwarz, J.M., and Lustig, R.H. Markus, C.R., Rogers, P.J., Brouns, F., and Schepers, R. (2017). Eating depen-
(2010). The role of fructose in the pathogenesis of NAFLD and the metabolic dence and weight gain; no human evidence for a ‘‘sugar-addiction’’ model of
syndrome. Nat. Rev. Gastroenterol. Hepatol 7, 251–264. overweight. Appetite 114, 64–72.

Lin, J., Puigserver, P., Donovan, J., Tarr, P., and Spiegelman, B.M. (2002). Marriott, B.P., Cole, N., and Lee, E. (2009). National estimates of dietary fruc-
Peroxisome proliferator-activated receptor gamma coactivator 1beta (PGC- tose intake increased from 1977 to 2004 in the United States. J. Nutr 139,
1beta), a novel PGC-1-related transcription coactivator associated with host 1228S–1235S.
cell factor. J. Biol. Chem. 277, 1645–1648.
Marthaler, T.M., and Froesch, E.R. (1967). Hereditary fructose intolerance.
Lin, J., Yang, R., Tarr, P.T., Wu, P.H., Handschin, C., Li, S., Yang, W., Pei, L., Dental status of eight patients. Br. Dent. J 123, 597–599.
Uldry, M., Tontonoz, P., et al. (2005a). Hyperlipidemic effects of dietary satu-
rated fats mediated through PGC-1beta coactivation of SREBP. Cell 120, Martini, A.C., Tissera, A., Estofán, D., Molina, R.I., Mangeaud, A., de Cuneo,
261–273. M.F., and Ruiz, R.D. (2010). Overweight and seminal quality: a study of 794 pa-
tients. Fertil. Steril 94, 1739–1743.
Lin, J., Handschin, C., and Spiegelman, B.M. (2005b). Metabolic control Mayes, P.A. (1993). Intermediary metabolism of fructose. Am. J. Clin. Nutr. 58
through the PGC-1 family of transcription coactivators. Cell Metab 1, 361–370. (Supplement ), 754S–765S.
Linden, A.G., Li, S., Choi, H.Y., Fang, F., Fukasawa, M., Uyeda, K., Hammer, McGarry, J.D. (2002). Banting lecture 2001: dysregulation of fatty acid meta-
R.E., Horton, J.D., Engelking, L.J., and Liang, G. (2018). Interplay between bolism in the etiology of type 2 diabetes. Diabetes 51, 7–18.
ChREBP and SREBP-1c coordinates postprandial glycolysis and lipogenesis
in livers of mice. J. Lipid Res. 59, 475–487. McKeown, N.M., Dashti, H.S., Ma, J., Haslam, D.E., Kiefte-de Jong, J.C.,
Smith, C.E., Tanaka, T., Graff, M., Lemaitre, R.N., Rybin, D., et al. (2018).
Liu, L., Li, T., Liao, Y., Wang, Y., Gao, Y., Hu, H., Huang, H., Wu, F., Chen, Y.- Sugar-sweetened beverage intake associations with fasting glucose and insu-
G., Xu, S., et al. (2020). Triose kinase controls the lipogenic potential of fruc- lin concentrations are not modified by selected genetic variants in a ChREBP-
tose and dietary tolerance. Cell Metab 32, 605–618.e7. FGF21 pathway: a meta-analysis. Diabetologia 61, 317–330.

López-Casillas, F., Bai, D.H., Luo, X.C., Kong, I.S., Hermodson, M.A., and Kim, Mela, D.J., and Woolner, E.M. (2018). Perspective: total, added, or free? What
K.H. (1988). Structure of the coding sequence and primary amino acid kind of sugars should we be talking about? Adv. Nutr 9, 63–69.
sequence of acetyl-coenzyme A carboxylase. Proc. Natl. Acad. Sci. USA 85,
5784–5788. Melchior, C., Gourcerol, G., Déchelotte, P., Leroi, A.M., and Ducrotté, P.
(2014). Symptomatic fructose malabsorption in irritable bowel syndrome: a
Lorenzi, M. (2007). The polyol pathway as a mechanism for diabetic retinop- prospective study. United European Gastroenterol. J. 2, 131–137.
athy: attractive, elusive, and resilient. Exp. Diabetes Res. 2007, 61038.
Middleton, R.J. (1990). Hexokinases and glucokinases. Biochem. Soc. Trans.
Luo, W., Hu, H., Chang, R., Zhong, J., Knabel, M., O’Meally, R., Cole, R.N., 18, 180–183.
Pandey, A., and Semenza, G.L. (2011). Pyruvate kinase M2 is a PHD3-stimu-
Mirtschink, P., Krishnan, J., Grimm, F., Sarre, A., Hörl, M., Kayikci, M., Fank-
lated coactivator for hypoxia-inducible factor 1. Cell 145, 732–744.
hauser, N., Christinat, Y., Cortijo, C., Feehan, O., et al. (2015). HIF-driven
SF3B1 induces KHK-C to enforce fructolysis and heart disease. Nature 522,
Lustig, R.H. (2010). Fructose: metabolic, hedonic, and societal parallels with
444–449.
ethanol. J. Am. Diet. Assoc. 110, 1307–1321.
Mock, D.M., Perman, J.A., Thaler, M., and Morris, R.C. (1983). Chronic fruc-
Ma, L., Robinson, L.N., and Towle, H.C. (2006). ChREBP*Mlx is the principal tose intoxication after infancy in children with hereditary fructose intolerance.
mediator of glucose-induced gene expression in the liver. J. Biol. Chem. A cause of growth retardation. N. Engl. J. Med 309, 764–770.
281, 28721–28730.
Mosca, A., Nobili, V., De Vito, R., Crudele, A., Scorletti, E., Villani, A., Alisi, A.,
Ma, J., Fox, C.S., Jacques, P.F., Speliotes, E.K., Hoffmann, U., Smith, C.E., and Byrne, C.D. (2017). Serum uric acid concentrations and fructose con-
Saltzman, E., and McKeown, N.M. (2015). Sugar-sweetened beverage, diet sumption are independently associated with NASH in children and adoles-
soda, and fatty liver disease in the Framingham Heart Study cohorts. cents. J. Hepatol 66, 1031–1036.
J. Hepatol 63, 462–469.
Nagai, Y., Yonemitsu, S., Erion, D.M., Iwasaki, T., Stark, R., Weismann, D.,
Ma, J., Jacques, P.F., Meigs, J.B., Fox, C.S., Rogers, G.T., Smith, C.E., Hruby, Dong, J., Zhang, D., Jurczak, M.J., Löffler, M.G., et al. (2009). The role of
A., Saltzman, E., and McKeown, N.M. (2016). Sugar-sweetened beverage but peroxisome proliferator-activated receptor gamma coactivator-1 beta in the
not diet soda consumption is positively associated with progression of insulin pathogenesis of fructose-induced insulin resistance. Cell Metab 9, 252–264.
resistance and prediabetes. J. Nutr 146, 2544–2550.
Nair, S.P., Chacko, V., Arnold, C., and Diehl, A.M. (2003). Hepatic ATP reserve
Macdonald, I. (1966). Influence of fructose and glucose on serum lipid levels in and efficiency of replenishing: comparison between obese and nonobese
men and pre- and postmenopausal women. Am. J. Clin. Nutr 18, 369–372. normal individuals. Am. J. Gastroenterol 98, 466–470.

Maersk, M., Belza, A., Stødkilde-Jørgensen, H., Ringgaard, S., Chabanova, E., Nelson, G., Hoon, M.A., Chandrashekar, J., Zhang, Y., Ryba, N.J., and Zuker,
Thomsen, H., Pedersen, S.B., Astrup, A., and Richelsen, B. (2012). Sucrose- C.S. (2001). Mammalian sweet taste receptors. Cell 106, 381–390.
sweetened beverages increase fat storage in the liver, muscle, and visceral
fat depot: a 6-mo randomized intervention study. Am. J. Clin. Nutr 95, Newgard, C.B., An, J., Bain, J.R., Muehlbauer, M.J., Stevens, R.D., Lien, L.F.,
283–289. Haqq, A.M., Shah, S.H., Arlotto, M., Slentz, C.A., et al. (2009). A branched-
chain amino acid-related metabolic signature that differentiates obese and
Malik, V.S., Popkin, B.M., Bray, G.A., Després, J.P., and Hu, F.B. (2010). lean humans and contributes to insulin resistance. Cell Metab 9, 311–326.
Sugar-sweetened beverages, obesity, type 2 diabetes mellitus, and cardio-
Niculescu, L., Veiga-da-Cunha, M., and Van Schaftingen, E. (1997). Investiga-
vascular disease risk. Circulation 121, 1356–1364.
tion on the mechanism by which fructose, hexitols and other compounds regu-
late the translocation of glucokinase in rat hepatocytes. Biochem. J. 321,
Malik, V.S., Pan, A., Willett, W.C., and Hu, F.B. (2013). Sugar-sweetened bev- 239–246.
erages and weight gain in children and adults: a systematic review and meta-
analysis. Am. J. Clin. Nutr 98, 1084–1102. Niewoehner, C.B., Gilboe, D.P., Nuttall, G.A., and Nuttall, F.Q. (1984). Meta-
bolic effects of oral fructose in the liver of fasted rats. Am. J. Physiol 247,
Manolescu, A.R., Witkowska, K., Kinnaird, A., Cessford, T., and Cheeseman, E505–E512.
C. (2007). Facilitated hexose transporters: new perspectives on form and func-
tion. Physiology (Bethesda) 22, 234–240. €, E.A., and Ojala, K. (1965). Induction of hyperglyceridemia by fructose
Nikkila
in the rat. Life Sci. 4, 937–943.
Marek, G., Pannu, V., Shanmugham, P., Pancione, B., Mascia, D., Crosson, S.,
Ishimoto, T., and Sautin, Y.Y. (2015). Adiponectin resistance and proinflamma- Nikolaou, K.C., Vatandaslar, H., Meyer, C., Schmid, M.W., Tuschl, T., and Stof-
tory changes in the visceral adipose tissue induced by fructose consumption fel, M. (2019). The RNA-binding protein A1CF regulates hepatic fructose and
via ketohexokinase-dependent pathway. Diabetes 64, 508–518. glycerol metabolism via alternative RNA splicing. Cell Rep 29, 283–300.e8.

2350 Cell Metabolism 33, December 7, 2021


ll
Review
Niwa, H., Iizuka, K., Kato, T., Wu, W., Tsuchida, H., Takao, K., Horikawa, Y., Pontzer, H., Raichlen, D.A., Wood, B.M., Mabulla, A.Z.P., Racette, S.B., and
and Takeda, J. (2018). ChREBP rather than SHP regulates hepatic VLDL Marlowe, F.W. (2012). Hunter-gatherer energetics and human obesity. PLoS
secretion. Nutrients 10, 321. One 7, e40503.

Oates, P.J. (2002). Polyol pathway and diabetic peripheral neuropathy. Int. Preston, G.M., and Calle, R.A. (2010). Elevated serum sorbitol and not fructose
Rev. Neurobiol. 50, 325–392. in type 2 diabetic patients. Biomark. Insights 5, 33–38.

Oberhaensli, R.D., Galloway, G.J., Taylor, D.J., Bore, P.J., and Radda, G.K. Queiroz-Leite, G.D., Crajoinas, R.O., Neri, E.A., Bezerra, C.N.A., Girardi,
(1986). Assessment of human liver metabolism by phosphorus-31 magnetic A.C.C., Rebouças, N.A., and Malnic, G. (2012). Fructose acutely stimulates
resonance spectroscopy. Br. J. Radiol 59, 695–699. NHE3 activity in kidney proximal tubule. Kidney Blood Press. Res. 36,
320–334.
Oberhaensli, R.D., Rajagopalan, B., Taylor, D.J., Radda, G.K., Collins, J.E.,
Leonard, J.V., Schwarz, H., and Herschkowitz, N. (1987). Study of hereditary Raichlen, D.A., Pontzer, H., Harris, J.A., Mabulla, A.Z.P., Marlowe, F.W., Josh
fructose intolerance by use of 31P magnetic resonance spectroscopy. Lancet Snodgrass, J., Eick, G., Colette Berbesque, J., Sancilio, A., and Wood, B.M.
2, 931–934. (2017). Physical activity patterns and biomarkers of cardiovascular disease
risk in hunter-gatherers. Am. J. Hum. Biol. 29. https://doi.org/10.1002/
Oh, A.R., Sohn, S., Lee, J., Park, J.M., Nam, K.T., Hahm, K.B., Kim, Y.B., Lee, ajhb.22919.
H.J., and Cha, J.Y. (2018). ChREBP deficiency leads to diarrhea-predominant
irritable bowel syndrome. Metabolism 85, 286–297. Raivio, K.O., Becker, 7 A., Meyer, L.J., Greene, M.L., Nuki, G., and Seegmiller,
J.E. (1975). Stimulation of human purine synthesis de novo by fructose infu-
Oppelt, S.A., Sennott, E.M., and Tolan, D.R. (2015). Aldolase-B knockout in sion. Metabolism 24, 861–869.
mice phenocopies hereditary fructose intolerance in humans. Mol. Genet.
Metab 114, 445–450. Rambaud, P., Joannard, A., Bost, M., Marchal, A., Rachail, M., and Roget, J.
(1973). Arch. Fr. Pediatr 30, 1051–1062.
Orozco, J.M., Krawczyk, P.A., Scaria, S.M., Cangelosi, A.L., Chan, S.H., Kun-
chok, T., Lewis, C.A., and Sabatini, D.M. (2020). Dihydroxyacetone phosphate Rand, E.B., Depaoli, A.M., Davidson, N.O., Bell, G.I., and Burant, C.F. (1993).
signals glucose availability to mTORC1. Nat. Metab 2, 893–901. Sequence, tissue distribution, and functional characterization of the rat fruc-
tose transporter GLUT5. Am. J. Physiol 264, G1169–G1176.
Ouyang, X., Cirillo, P., Sautin, Y., McCall, S., Bruchette, J.L., Diehl, A.M., John-
son, R.J., and Abdelmalek, M.F. (2008). Fructose consumption as a risk factor Rao, S.S.C., Attaluri, A., Anderson, L., and Stumbo, P. (2007). Ability of the
for non-alcoholic fatty liver disease. J. Hepatol 48, 993–999. normal human small intestine to absorb fructose: evaluation by breath testing.
Clin. Gastroenterol. Hepatol 5, 959–963.
Pagliassotti, M.J., and Prach, P.A. (1995). Quantity of sucrose alters the tissue
pattern and time course of insulin resistance in young rats. Am. J. Physiol 269, Reaven, G.M. (1988). Banting lecture 1988. Role of insulin resistance in human
R641–R646. disease. Diabetes 37, 1595–1607.

Pagliassotti, M.J., Shahrokhi, K.A., and Moscarello, M. (1994). Involvement of Reaven, G.M., Risser, T.R., Chen, Y.D., and Reaven, E.P. (1979). Characteriza-
liver and skeletal muscle in sucrose-induced insulin resistance: dose- tion of a model of dietary-induced hypertriglyceridemia in young, nonobese
response studies. Am. J. Physiol. 266, R1637–R1644. rats. J. Lipid Res. 20, 371–378.

Patel, C., Douard, V., Yu, S., Gao, N., and Ferraris, R.P. (2015a). Transport, Redondo, M., Hernández-Aguado, I., and Lumbreras, B. (2018). The impact of
metabolism, and endosomal trafficking-dependent regulation of intestinal the tax on sweetened beverages: a systematic review. Am. J. Clin. Nutr 108,
fructose absorption. FASEB J 29, 4046–4058. 548–563.

Patel, C., Douard, V., Yu, S., Tharabenjasin, P., Gao, N., and Ferraris, R.P. Reinicke, K., Sotomayor, P., Cisterna, P., Delgado, C., Nualart, F., and Godoy,
(2015b). Fructose-induced increases in expression of intestinal fructolytic A. (2012). Cellular distribution of Glut-1 and Glut-5 in benign and malignant hu-
and gluconeogenic genes are regulated by GLUT5 and KHK. Am. J. Physiol. man prostate tissue. J. Cell. Biochem 113, 553–562.
Regul. Integr. Comp. Physiol 309, R499–R509.
Reiser, S., and Hallfrisch, J. (1977). Insulin sensitivity and adipose tissue
Patel, C., Sugimoto, K., Douard, V., Shah, A., Inui, H., Yamanouchi, T., and weight of rats fed starch or sucrose diets ad libitum or in meals. J. Nutr 107,
Ferraris, R.P. (2015c). Effect of dietary fructose on portal and systemic serum 147–155.
fructose levels in rats and in KHK-/- and GLUT5-/- mice. Am. J. Physiol. Gas-
trointest. Liver Physiol. 309, G779–G790. Rosinger, A., Herrick, K., Gahche, J., and Park, S. (2017). Sugar-sweetened
beverage consumption among U.S. adults, 2011–2014. NCHS Data Brief
Pedersen, K.B., Zhang, P., Doumen, C., Charbonnet, M., Lu, D., Newgard, 270, 1–8.
C.B., Haycock, J.W., Lange, A.J., and Scott, D.K. (2007). The promoter for
the gene encoding the catalytic subunit of rat glucose-6-phosphatase con- Santer, R., Hillebrand, G., Steinmann, B., and Schaub, J. (2003). Intestinal
tains two distinct glucose-responsive regions. Am. J. Physiol. Endocrinol. glucose transport: evidence for a membrane traffic–based pathway in hu-
Metab. 292, E788–E801. mans. Gastroenterology 124, 34–39.

€, E.A. (1972). Hereditary fructose


Perheentupa, J., Raivio, K.O., and Nikkila Schmidt, S., Joost, H.G., and Schu €rmann, A. (2009). GLUT8, the enigmatic
intolerance. Acta Med. Scand. Suppl. 542, 65–75. intracellular hexose transporter. Am. J. Physiol. Endocrinol. Metab. 296,
E614–E618.
Petersen, M.C., and Shulman, G.I. (2018). Mechanisms of insulin action and in-
sulin resistance. Physiol. Rev. 98, 2133–2223. Schwarz, J.M., Noworolski, S.M., Wen, M.J., Dyachenko, A., Prior, J.L., Wein-
berg, M.E., Herraiz, L.A., Tai, V.W., Bergeron, N., Bersot, T.P., et al. (2015). Ef-
Petersen, A., Steinmann, B., and Gitzelmann, R. (1992). Essential fructosuria: fect of a high-fructose weight-maintaining diet on lipogenesis and liver fat.
increased levels of fructose 3-phosphate in erythrocytes. Enzyme 46, J. Clin. Endocrinol. Metab 100, 2434–2442.
319–323.
Schwarz, J.M., Noworolski, S.M., Erkin-Cakmak, A., Korn, N.J., Wen, M.J.,
Petersen, K.F., Laurent, D., Yu, C., Cline, G.W., and Shulman, G.I. (2001). Stim- Tai, V.W., Jones, G.M., Palii, S.P., Velasco-Alin, M., Pan, K., et al. (2017). Ef-
ulating effects of low-dose fructose on insulin-stimulated hepatic glycogen fects of dietary fructose restriction on liver fat, de novo lipogenesis, and insulin
synthesis in humans. Diabetes 50, 1263–1268. kinetics in children with obesity. Gastroenterology 153, 743–752.

Pfinder, M., Heise, T.L., Hilton Boon, M., Pega, F., Fenton, C., Griebler, U., Gar- Schwimmer, J.B., Ugalde-Nicalo, P., Welsh, J.A., Angeles, J.E., Cordero, M.,
tlehner, G., Sommer, I., Katikireddi, S.V., and Lhachimi, S.K. (2020). Taxation Harlow, K.E., Alazraki, A., Durelle, J., Knight-Scott, J., Newton, K.P., et al.
of unprocessed sugar or sugar-added foods for reducing their consumption (2019). Effect of a low free sugar diet vs usual diet on nonalcoholic fatty liver
and preventing obesity or other adverse health outcomes. Cochrane Database disease in adolescent boys: a randomized clinical trial. JAMA 321, 256–265.
Syst. Rev. 4, CD012333.
Sclafani, A., and Ackroff, K. (2021). Fructose appetition in ‘‘taste-blind’’ P2X2/
Pollard-Knight, D., and Cornish-Bowden, A. (1982). Mechanism of liver gluco- P2X3 double knockout mice. bioRxiv. https://doi.org/10.1101/2021.07.22.
kinase. Mol. Cell. Biochem 44, 71–80. 453428.

Cell Metabolism 33, December 7, 2021 2351


ll
Review
Sclafani, A., Marambaud, P., and Ackroff, K. (2014). Sucrose-conditioned fla- Stanhope, K.L., Bremer, A.A., Medici, V., Nakajima, K., Ito, Y., Nakano, T.,
vor preferences in sweet ageusic T1r3 and Calhm1 knockout mice. Physiol. Chen, G., Fong, T.H., Lee, V., Menorca, R.I., et al. (2011). Consumption of fruc-
Behav 126, 25–29. tose and high fructose corn syrup increase postprandial triglycerides,
LDL-cholesterol, and apolipoprotein-B in young men and women. J. Clin. En-
Shah, A., Dagdeviren, S., Lewandowski, J.P., Schmider, A.B., Ricci-Blair, docrinol. Metab. 96, E1596–E1605.
E.M., Natarajan, N., Hundal, H., Noh, H.L., Friedline, R.H., Vidoudez, C.,
et al. (2020). Thioredoxin interacting protein is required for a chronic energy- Stanhope, K.L., Medici, V., Bremer, A.A., Lee, V., Lam, H.D., Nunez, M.V.,
rich diet to promote intestinal fructose absorption. iScience 23, 101521. Chen, G.X., Keim, N.L., and Havel, P.J. (2015). A dose-response study of
consuming high-fructose corn syrup-sweetened beverages on lipid/lipopro-
Shapiro, A., Mu, W., Roncal, C., Cheng, K.Y., Johnson, R.J., and Scarpace, tein risk factors for cardiovascular disease in young adults. Am. J. Clin. Nutr
P.J. (2008). Fructose-induced leptin resistance exacerbates weight gain in 101, 1144–1154.
response to subsequent high-fat feeding. Am. J. Physiol. Regul. Integr.
Comp. Physiol 295, R1370–R1375. Steinmann, B., Baerlocher, K., and Gitzelmann, R. (1975). Nutr. Metab. 18
(suppl 1), 115–132.
Sharpton, S.R., Schnabl, B., Knight, R., and Loomba, R. (2021). Current con-
cepts, opportunities, and challenges of gut microbiome-based personalized Steinmann, B., Gitzelmann, R., and Van Den Berghe, G. (2019). Disorders of
medicine in nonalcoholic fatty liver disease. Cell Metab 33, 21–32. fructose metabolism. In The Online Metabolic and Molecular Bases of In-
herited Disease, D.L. Valle, S. Antonarakis, A. Ballabio, A.L. Beaudet, and
Shi, J.-H., Lu, J.-Y., Chen, H.-Y., Wei, C.-C., Xu, X., Li, H., Bai, Q., Xia, F.-Z., G.A. Mitchell, eds. (McGraw Hill).
Lam, S.M., Zhang, H., et al. (2020). Liver ChREBP protects against fructose-
induced glycogenic hepatotoxicity by regulating L-type pyruvate kinase. Dia- Stu€mpel, F., Burcelin, R., Jungermann, K., and Thorens, B. (2001). Normal ki-
betes 69, 591–602. netics of intestinal glucose absorption in the absence of GLUT2: evidence for a
transport pathway requiring glucose phosphorylation and transfer into the
Shimano, H., and Sato, R. (2017). SREBP-regulated lipid metabolism: conver- endoplasmic reticulum. Proc. Natl. Acad. Sci. USA 98, 11330–11335.
gent physiology - divergent pathophysiology. Nat. Rev. Endocrinol 13,
710–730. Sul, H.S., Wise, L.S., Brown, M.L., and Rubin, C.S. (1984). Cloning of cDNA se-
quences for murine ATP-citrate lyase. Construction of recombinant plasmids
Sillero, M.A., Sillero, A., and Sols, A. (1969). Enzymes involved in fructose using an immunopurified mRNA template and evidence for the nutritional regu-
metabolism in lir and the glyceraldehyde metabolic crossroads. Eur. J. Bio- lation of ATP-citrate lyase mRNA content in mouse liver. J. Biol. Chem. 259,
chem 10, 345–350. 1201–1205.
Simons, N., Debray, F.G., Schaper, N.C., Kooi, M.E., Feskens, E.J.M., Hollak, Sullivan, J.S., Le, M.T., Pan, Z., Rivard, C., Love-Osborne, K., Robbins, K.,
C.E.M., Lindeboom, L., Koek, G.H., Bons, J.A.P., Lefeber, D.J., et al. (2019). Johnson, R.J., Sokol, R.J., and Sundaram, S.S. (2015). Oral fructose absorp-
Patients with aldolase B deficiency are characterized by increased intrahe- tion in obese children with non-alcoholic fatty liver disease. Pediatr. Obes 10,
patic triglyceride content. J. Clin. Endocrinol. Metab 104, 5056–5064. 188–195.

Simons, N., Veeraiah, P., Simons, P.I.H.G., Schaper, N.C., Kooi, M.E., Schrau- Sun, R.Q., Wang, H., Zeng, X.Y., Chan, S.M.H., Li, S.P., Jo, E., Leung, S.L.,
wen-Hinderling, V.B., Feskens, E.J.M., van der Ploeg, E.M.C.L., Van den Molero, J.C., and Ye, J.M. (2015). IRE1 impairs insulin signaling transduction
Eynde, M.D.G., Schalkwijk, C.G., et al. (2021). Effects of fructose restriction of fructose-fed mice via JNK independent of excess lipid. Biochim. Biophys.
on liver steatosis (FRUITLESS); a double-blind randomized controlled trial. Acta 1852, 156–165.
Am. J. Clin. Nutr 113, 391–400.
Talukdar, S., Owen, B.M., Song, P., Hernandez, G., Zhang, Y., Zhou, Y., Scott,
Singh, A.K., Amlal, H., Haas, P.J., Dringenberg, U., Fussell, S., Barone, S.L., W.T., Paratala, B., Turner, T., Smith, A., et al. (2016). FGF21 regulates sweet
Engelhardt, R., Zuo, J., Seidler, U., and Soleimani, M. (2008). Fructose- and alcohol preference. Cell Metab 23, 344–349.
induced hypertension: essential role of chloride and fructose absorbing trans-
porters PAT1 and Glut5. Kidney Int 74, 438–447. Tan, H.E., Sisti, A.C., Jin, H., Vignovich, M., Villavicencio, M., Tsang, K.S.,
Goffer, Y., and Zuker, C.S. (2020). The gut-brain axis mediates sugar prefer-
Sleder, J., Chen, Y.D., Cully, M.D., and Reaven, G.M. (1980). Hyperinsulinemia ence. Nature 580, 511–516.
in fructose-induced hypertriglyceridemia in the rat. Metabolism 29, 303–305.
Tanaka, T., Ngwa, J.S., van Rooij, F.J.A., Zillikens, M.C., Wojczynski, M.K.,
Smith, G.I., Shankaran, M., Yoshino, M., Schweitzer, G.G., Chondronikola, M., Frazier-Wood, A.C., Houston, D.K., Kanoni, S., Lemaitre, R.N., Luan, J.,
Beals, J.W., Okunade, A.L., Patterson, B.W., Nyangau, E., Field, T., et al. et al. (2013). Genome-wide meta-analysis of observational studies shows
(2020). Insulin resistance drives hepatic de novo lipogenesis in nonalcoholic common genetic variants associated with macronutrient intake. Am. J. Clin.
fatty liver disease. J. Clin. Invest 130, 1453–1460. Nutr 97, 1395–1402.
Softic, S., Gupta, M.K., Wang, G.-X., Fujisaka, S., O’Neill, B.T., Rao, T.N., Wil- Tappy, L. (2018). Fructose-containing caloric sweeteners as a cause of obesity
loughby, J., Harbison, C., Fitzgerald, K., Ilkayeva, O., et al. (2018). Divergent and metabolic disorders. J. Exp. Biol. 221.
effects of glucose and fructose on hepatic lipogenesis and insulin signaling.
J. Clin. Invest 128, 1199. Tappy, L., and Rosset, R. (2017). Fructose metabolism from a functional
perspective: implications for athletes. Sports Med 47, 23–32.
Softic, S., Meyer, J.G., Wang, G.X., Gupta, M.K., Batista, T.M., Lauritzen,
H.P.M.M., Fujisaka, S., Serra, D., Herrero, L., Willoughby, J., et al. (2019). Di- Taskinen, M.R., Söderlund, S., Bogl, L.H., Hakkarainen, A., Matikainen, N.,
etary sugars alter hepatic fatty acid oxidation via transcriptional and post- €inen, K.H., Ra
Pietila € sa
€nen, S., Lundbom, N., Björnson, E., Eliasson, B., et al.
translational modifications of mitochondrial proteins. Cell Metab 30, (2017). Adverse effects of fructose on cardiometabolic risk factors and hepatic
735–753.e4. lipid metabolism in subjects with abdominal obesity. J. Intern. Med 282,
187–201.
Softic, S., Stanhope, K.L., Boucher, J., Divanovic, S., Lanaspa, M.A., Johnson,
R.J., and Kahn, C.R. (2020). Fructose and hepatic insulin resistance. Crit. Rev. Taylor, S.R., Ramsamooj, S., Liang, R.J., Katti, A., Pozovskiy, R., Vasan, N.,
Clin. Lab. Sci. 57, 308–322. Hwang, S.K., Nahiyaan, N., Francoeur, N.J., Schatoff, E.M., et al. (2021). Die-
tary fructose improves intestinal cell survival and nutrient absorption. Nature
Stanhope, K.L. (2016). Sugar consumption, metabolic disease and obesity: the 597, 263–267.
state of the controversy. Crit. Rev. Clin. Lab. Sci. 53, 52–67.
Teff, K.L., Elliott, S.S., Tschöp, M., Kieffer, T.J., Rader, D., Heiman, M., Town-
Stanhope, K.L., Griffen, S.C., Bair, B.R., Swarbrick, M.M., Keim, N.L., and Ha- send, R.R., Keim, N.L., D’Alessio, D., and Havel, P.J. (2004). Dietary fructose
vel, P.J. (2008). Twenty-four-hour endocrine and metabolic profiles following reduces circulating insulin and leptin, attenuates postprandial suppression
consumption of high-fructose corn syrup-, sucrose-, fructose-, and glucose- of ghrelin, and increases triglycerides in women. J. Clin. Endocrinol. Metab
sweetened beverages with meals. Am. J. Clin. Nutr 87, 1194–1203. 89, 2963–2972.

Stanhope, K.L., Schwarz, J.M., Keim, N.L., Griffen, S.C., Bremer, A.A., Gra- Ter Horst, K.W., and Serlie, M.J. (2017). Fructose consumption, lipogenesis,
ham, J.L., Hatcher, B., Cox, C.L., Dyachenko, A., Zhang, W., et al. (2009). and non-alcoholic fatty liver disease. Nutrients 9, 981.
Consuming fructose-sweetened, not glucose-sweetened, beverages in-
creases visceral adiposity and lipids and decreases insulin sensitivity in over- Ter Horst, K.W., Schene, M.R., Holman, R., Romijn, J.A., and Serlie, M.J.
weight/obese humans. J. Clin. Invest 119, 1322–1334. (2016). Effect of fructose consumption on insulin sensitivity in nondiabetic

2352 Cell Metabolism 33, December 7, 2021


ll
Review
subjects: a systematic review and meta-analysis of diet-intervention trials. Am. sweetened beverage consumption among 0-year to 5-year olds. Obes. Rev.
J. Clin. Nutr 104, 1562–1576. 19, 1504–1524.

Ter Horst, K.W., Vatner, D.F., Zhang, D., Cline, G.W., Ackermans, M.T., Ne- Virani, S.S., Alonso, A., Benjamin, E.J., Bittencourt, M.S., Callaway, C.W., Car-
derveen, A.J., Verheij, J., Demirkiran, A., van Wagensveld, B.A., Dallinga- son, A.P., Chamberlain, A.M., Chang, A.R., Cheng, S., Delling, F.N., et al.
Thie, G.M., et al. (2020). Hepatic insulin resistance is not pathway selective (2020). Heart disease and stroke statistics-2020 update: a report from the
in humans with nonalcoholic fatty liver disease. Diabetes Care 44, 489–498. American Heart Association. Circulation 141, e139–e596.

Thresher, J.S., Podolin, D.A., Wei, Y., Mazzeo, R.S., and Pagliassotti, M.J. Volynets, V., Machann, J., Ku€per, M.A., Maier, I.B., Spruss, A., Königsrainer,
(2000). Comparison of the effects of sucrose and fructose on insulin action A., Bischoff, S.C., and Bergheim, I. (2013). A moderate weight reduction
and glucose tolerance. Am. J. Physiol. Regul. Integr. Comp. Physiol. 279, through dietary intervention decreases hepatic fat content in patients with
R1334–R1340. non-alcoholic fatty liver disease (NAFLD): a pilot study. Eur. J. Nutr 52,
527–535.
Thurston, J.H., Jones, E.M., and Hauhart, R.E. (1974). Decrease and inhibition
of liver glycogen phosphorylase after fructose. An experimental model for the von Holstein-Rathlou, S., Bondurant, L.D., Peltekian, L., Naber, M.C., Yin, T.C.,
study of hereditary fructose intolerance. Diabetes 23, 597–604. Claflin, K.E., Urizar, A.I., Madsen, A.N., Ratner, C., Holst, B., et al. (2016).
FGF21 mediates endocrine control of simple sugar intake and sweet taste
Tin, A., Marten, J., Halperin Kuhns, V.L., Li, Y., Wuttke, M., Kirsten, H., Sieber, preference by the liver. Cell Metab 23, 335–343.
K.B., Qiu, C., Gorski, M., Yu, Z., et al. (2019). Target genes, variants, tissues
and transcriptional pathways influencing human serum urate levels. Nat. von Philipsborn, P., Stratil, J.M., Burns, J., Busert, L.K., Pfadenhauer, L.M.,
Genet. 51, 1459–1474. Polus, S., Holzapfel, C., Hauner, H., and Rehfuess, E.A. (2020). Environmental
interventions to reduce the consumption of sugar-sweetened beverages:
Tobey, T.A., Greenfield, M., Kraemer, F., and Reaven, G.M. (1981). Relation- abridged cochrane systematic review. Obes. Facts 13, 397–417.
ship between insulin resistance, insulin secretion, very low density lipoprotein
kinetics, and plasma triglyceride levels in normotriglyceridemic man. Meta- Wakil, S.J., Stoops, J.K., and Joshi, V.C. (1983). Fatty acid synthesis and its
bolism 30, 165–171. regulation. Annu. Rev. Biochem 52, 537–579.
Todoric, J., Di Caro, G., Reibe, S., Henstridge, D.C., Green, C.R., Vrbanac, A., Walker, R.W., Lê, K.A., Davis, J., Alderete, T.L., Cherry, R., Lebel, S., and
Ceteci, F., Conche, C., McNulty, R., Shalapour, S., et al. (2020). Fructose stim- Goran, M.I. (2012). High rates of fructose malabsorption are associated with
ulated de novo lipogenesis is promoted by inflammation. Nat. Metab 2, reduced liver fat in obese African Americans. J. Am. Coll. Nutr 31, 369–374.
1034–1045.
Wang, L., Gillis-Smith, S., Peng, Y., Zhang, J., Chen, X., Salzman, C.D., Ryba,
Togo, J., Hu, S., Li, M., Niu, C., and Speakman, J.R. (2019). Impact of dietary N.J.P., and Zuker, C.S. (2018). The coding of valence and identity in the
sucrose on adiposity and glucose homeostasis in C57BL/6J mice depends on mammalian taste system. Nature 558, 127–131.
mode of ingestion: liquid or solid. Mol. Metab 27, 22–32.
Wasserman, D., Hoekstra, J.H., Tolia, V., Taylor, C.J., Kirschner, B.S., Takeda,
Topping, D.L., and Mayes, P.A. (1971). The concentration of fructose, glucose J., Bell, G.I., Taub, R., and Rand, E.B. (1996). Molecular analysis of the fructose
and lactate in the splanchnic blood vessels of rats absorbing fructose. Nutr. transporter gene (GLUT5) in isolated fructose malabsorption. J. Clin. Invest 98,
Metab 13, 331–338. 2398–2402.
Topping, D.L., and Mayes, P.A. (1972). The immediate effects of insulin and
Wei, Y., and Pagliassotti, M.J. (2004). Hepatospecific effects of fructose on c-
fructose on the metabolism of the perfused liver. Changes in lipoprotein secre-
jun NH2-terminal kinase: implications for hepatic insulin resistance. Am. J.
tion, fatty acid oxidation and esterification, lipogenesis and carbohydrate
Physiol. Endocrinol. Metab. 287, E926–E933.
metabolism. Biochem. J. 126, 295–311.
Wei, Y., Bizeau, M.E., and Pagliassotti, M.J. (2004). An acute increase in fruc-
Topping, D.L., and Mayes, P.A. (1976). Comparative effects of fructose and
tose concentration increases hepatic glucose-6-phosphatase mRNA via
glucose on the lipid and carbohydrate metabolism of perfused rat liver. Br.
mechanisms that are independent of glycogen synthase kinase-3 in rats.
J. Nutr 36, 113–126.
J. Nutr 134, 545–551.
Touger-Decker, R., and van Loveren, C. (2003). Sugars and dental caries. Am.
Wei, Y., Wang, D., and Pagliassotti, M.J. (2005). Fructose selectively modu-
J. Clin. Nutr 78, 881S–892S.
lates c-jun N-terminal kinase activity and insulin signaling in rat primary hepa-
Traussnigg, S., Kienbacher, C., Gajdosı́k, M., Valkovic
, L., Halilbasic, E., Stift, tocytes. J. Nutr 135, 1642–1646.
J., Rechling, C., Hofer, H., Steindl-Munda, P., Ferenci, P., et al. (2017). Ultra-
high-field magnetic resonance spectroscopy in non-alcoholic fatty liver dis- White, P.J., and Newgard, C.B. (2019). Branched-chain amino acids in dis-
ease: novel mechanistic and diagnostic insights of energy metabolism in ease. Science 363, 582–583.
non-alcoholic steatohepatitis and advanced fibrosis. Liver Int 37, 1544–1553.
White, P.J., McGarrah, R.W., Grimsrud, P.A., Tso, S.C., Yang, W.H., Halde-
U.S. Department of Agriculture; U.S. Department of Health and Human Services man, J.M., Grenier-Larouche, T., An, J., Lapworth, A.L., Astapova, I., et al.
(2020). Dietary guidelines for Americans. (2020–2025), (9th edition). https:// (2018). The BCKDH kinase and phosphatase integrate BCAA and lipid meta-
www.dietaryguidelines.gov/sites/default/files/2021-03/Dietary_Guidelines_ bolism via regulation of ATP-citrate lyase. Cell Metab 27, 1281–1293.e7.
for_Americans-2020-2025.pdf.
Woods, H.F., Eggleston, L.V., and Krebs, H.A. (1970). The cause of hepatic
Ushijima, K., Riby, J.E., and Kretchmer, N. (1995). Carbohydrate malabsorp- accumulation of fructose 1-phosphate on fructose loading. Biochem. J. 119,
tion. Pediatr. Clin. North Am. 42, 899–915. 501–510.

Uyeda, K., and Repa, J.J. (2006). Carbohydrate response element binding pro- World Health Organization (2015). Guideline: sugars intake for adults and chil-
tein, ChREBP, a transcription factor coupling hepatic glucose utilization and dren. https://www.who.int/publications-detail-redirect/9789241549028.
lipid synthesis. Cell Metab 4, 107–110.
Xi, B., Huang, Y., Reilly, K.H., Li, S., Zheng, R., Barrio-Lopez, M.T., Martinez-
Van Den Berghe, G., Hue, L., and Hers, H.G. (1973). Effect of administration of Gonzalez, M.A., and Zhou, D. (2015). Sugar-sweetened beverages and risk of
the fructose on the glycogenolytic action of glucagon. An investigation of the hypertension and CVD: a dose-response meta-analysis. Br. J. Nutr 113,
pathogeny of hereditary fructose intolerance. Biochem. J. 134, 637–645. 709–717.

van den Berghe, G., Bronfman, M., Vanneste, R., and Hers, H.G. (1977). The Yamashita, H., Takenoshita, M., Sakurai, M., Bruick, R.K., Henzel, W.J., Shil-
mechanism of adenosine triphosphate depletion in the liver after a load of fruc- linglaw, W., Arnot, D., and Uyeda, K. (2001). A glucose-responsive transcrip-
tose. A kinetic study of liver adenylate deaminase. Biochem. J. 162, 601–609. tion factor that regulates carbohydrate metabolism in the liver. Proc. Natl.
Acad. Sci. USA 98, 9116–9121.
Van Schaftingen, E. (1994). Short-term regulation of glucokinase. Diabetologia
37 (suppl 2), S43–S47. Yang, Q., Köttgen, A., Dehghan, A., Smith, A.V., Glazer, N.L., Chen, M.H.,
Chasman, D.I., Aspelund, T., Eiriksdottir, G., Harris, T.B., et al. (2010). Multiple
Vercammen, K.A., Frelier, J.M., Lowery, C.M., McGlone, M.E., Ebbeling, C.B., genetic loci influence serum urate levels and their relationship with gout and
and Bleich, S.N. (2018). A systematic review of strategies to reduce sugar- cardiovascular disease risk factors. Circ. Cardiovasc. Genet. 3, 523–530.

Cell Metabolism 33, December 7, 2021 2353


ll
Review
Yang, N., Hong, N.J., and Garvin, J.L. (2020). Dietary fructose enhances angio- Zavaroni, I., Chen, Y.I., Mondon, C.E., and Reaven, G.M. (1981). Ability of ex-
tensin II-stimulated Na+ transport via activation of PKC-a in renal proximal tu- ercise to inhibit carbohydrate-induced hypertriglyceridemia in rats. Meta-
bules. Am. J. Physiol. Renal Physiol. 318, F1513–F1519. bolism 30, 476–480.

Yoneyama, N., Crabbe, J.C., Ford, M.M., Murillo, A., and Finn, D.A. (2008).
Voluntary ethanol consumption in 22 inbred mouse strains. Alcohol 42, Zhang, X., Li, X., Liu, L., Hong, F., Zhao, H., Chen, L., Zhang, J., Jiang, Y.,
149–160. Zhang, J., and Luo, P. (2020). Dose-response association between sugar-
and artificially sweetened beverage consumption and the risk of metabolic
Yoo, T.W., Sung, K.C., Shin, H.S., Kim, B.J., Kim, B.S., Kang, J.H., Lee, M.H., syndrome: a meta-analysis of population-based epidemiological studies. Pub-
Park, J.R., Kim, H., Rhee, E.J., et al. (2005). Relationship between serum uric lic Health Nutr 24, 3892–3904.
acid concentration and insulin resistance and metabolic syndrome. Circ. J. 69,
928–933.
Zhao, G.Q., Zhang, Y., Hoon, M.A., Chandrashekar, J., Erlenbach, I., Ryba,
Yoshida, M., McKeown, N.M., Rogers, G., Meigs, J.B., Saltzman, E., D’Agos- N.J.P., and Zuker, C.S. (2003). The receptors for mammalian sweet and umami
tino, R., and Jacques, P.F. (2007). Surrogate markers of insulin resistance are taste. Cell 115, 255–266.
associated with consumption of sugar-sweetened drinks and fruit juice in mid-
dle and older-aged adults. J. Nutr 137, 2121–2127.
Zhang, D., Tong, X., VanDommelen, K., Gupta, N., Stamper, K., Brady, G.F.,
Yoshii, H., Uchino, H., Ohmura, C., Watanabe, K., Tanaka, Y., and Kawamori, Meng, Z., Lin, J., Rui, L., Omary, M.B., et al. (2017). Lipogenic transcription fac-
R. (2001). Clinical usefulness of measuring urinary polyol excretion by gas- tor ChREBP mediates fructose-induced metabolic adaptations to prevent
chromatography/mass-spectrometry in type 2 diabetes to assess polyol hepatotoxicity. J. Clin. Invest. 127, 2855–2867.
pathway activity. Diabetes Res. Clin. Pract 51, 115–123.

Zagorsky, J.L., and Smith, P.K. (2020). Who drinks soda pop? Economic sta- Zhao, S., Jang, C., Liu, J., Uehara, K., Gilbert, M., Izzo, L., Zeng, X., Trefely, S.,
tus and adult consumption of sugar-sweetened beverages. Econ. Hum. Biol. Fernandez, S., Carrer, A., et al. (2020). Dietary fructose feeds hepatic lipogen-
38, 100888. esis via microbiota-derived acetate. Nature 579, 586–591.

2354 Cell Metabolism 33, December 7, 2021

You might also like