Você está na página 1de 15

REVIEWS

Short-chain fatty acids in control of body


weight and insulin sensitivity
Emanuel E. Canfora, Johan W. Jocken and Ellen E. Blaak
Abstract | The connection between the gut microbiota and the aetiology of obesity and cardiometabolic disorders
is increasingly being recognized by clinicians. Our gut microbiota might affect the cardiometabolic phenotype
by fermenting indigestible dietary components and thereby producing short-chain fatty acids (SCFA). These
SCFA are not only of importance in gut health and as signalling molecules, but might also enter the systemic
circulation and directly affect metabolism or the function of peripheral tissues. In this Review, we discuss
the effects of three SCFA (acetate, propionate and butyrate) on energy homeostasis and metabolism, as well
as how these SCFA can beneficially modulate adipose tissue, skeletal muscle and liver tissue function. As a
result, these SCFA contribute to improved glucose homeostasis and insulin sensitivity. Furthermore, we also
summarize the increasing evidence for a potential role of SCFA as metabolic targets to prevent and counteract
obesity and its associated disorders in glucose metabolism and insulin resistance. However, most data are
derived from animal and in vitro studies, and consequently the importance of SCFA and differential SCFA
availability in human energy and substrate metabolism remains to be fully established. Well-controlled human
intervention studies investigating the role of SCFA on cardiometabolic health are, therefore, eagerly awaited.

Canfora, E. E. et al. Nat Rev. Endocrinol. advance online publication 11 August 2015; doi:10.1038/nrendo.2015.128

Introduction
Obesity is becoming one of the major health-care prob- of SCFA including acetate, butyrate and propionate.10
lems of the twenty-first century and is associated with a These SCFA have key functional roles in the patho-
wide spectrum of pathological disturbances in metabolic physiology of obesity and related disorders by affecting
organs predisposed towards insulin resistance, as well control of body weight via regulation of energy intake,
as type 2 diabetes mellitus (T2DM) and cardiovascular energy harvesting and host energy and substrate metab­
disease.1–3 A tight interplay between different insulin- olism.11–17 In addition, several mechanisms that link
sensitive organs regulates energy and substrate metab­ SCFA to insulin sensitivity and the development of
olism in the human body. Functional disturbances in T2DM have been proposed, including the inter­organ
these organs can, therefore, theoretically cause or con- effects on adipose tissue function and lipid storage
tribute to the development of disturbances in glucose capacity, inflammatory profile, and liver and skeletal
metabolism, insulin resistance and cardiometabolic muscle substrate metabolism (Figure 1).16,18,19
disease. A growing body of evidence suggests that our In this Review, we discuss the current knowledge
gut and its resident microbiota have a crucial role in the regarding the role of the main gut-derived SCFA—acetate,
regulation of energy and substrate metabolism and propionate and butyrate. We provide an overview of
the development of cardiometabolic diseases in rodent luminal SCFA production, metabolism and absorp-
models4–7 and humans.8,9 In humans with obesity, 7 days tion rates and also discuss the putative effects of SCFA
of treatment with vancomycin, an antibiotic against on energy homeostasis and control of body weight.
Department of Human Gram-positive bacteria, can modulate the composition The effect of SCFA and nondigestible carbohydrates on
Biology, NUTRIM School
for Nutrition and
of the gut microbiota and decrease peripheral insulin adipose tissue, skeletal muscle and liver tissue metab­
Translational Research sensitivity.8 Furthermore, transfer of intestinal micro­ olism and function in relation to insulin sensitivity will
in Metabolism, biota from lean donors to individuals with the metabolic also be discussed.
Maastricht University
Medical Center, syndrome can improve insulin sensitivity and increase
Universiteitssingel 50, the number of intestinal butyrate-producing microbiota Production and absorption of SCFA
6229 ER, Maastricht,
PO Box 616,
and faecal short-chain fatty acids (SCFA).9 SCFA in the human gut
6200 MD, Maastricht, Our gut microbiota ferment dietary components SCFA are end products of the intestinal microbial fer-
Netherlands (E.E.C., that are incompletely hydrolysed due to a lack of the mentation of indigestible foods; complex carbohydrates,
J.W.J., E.E.B.).
appropriate enzymes, which results in the production such as resistant starch or dietary fibre, are the main
Correspondence to: substrates for SCFA production.20 The most abundant
E.E.B.
e.blaak@ Competing interests SCFA in the gut are acetate, propionate and butyrate
maastrichtuniversity.nl The authors declare no competing interests. (which constitute >95% of the SCFA content), while

NATURE REVIEWS | ENDOCRINOLOGY ADVANCE ONLINE PUBLICATION | 1


© 2015 Macmillan Publishers Limited. All rights reserved
REVIEWS

Key points SCFA production.10,35–39 Furthermore, differences in sub-


strate availability, due to altered gut transit times could
■■ Short-chain fatty acids (SCFA), which are derived from gut microbial
shift microbial composition and concen­trations, which
fermentation of indigestible foods, have important metabolic functions and are
crucial for intestinal health
might lead to alterations in microbial SCFA produc-
■■ The discovery of SCFA receptors in many different tissues highlights that SCFA tion.10,35–39 Consistent with this observation, a cisapride-
are involved in the crosstalk between the gut and peripheral tissues induced reduction in gut transit time in humans was
■■ In addition to their role in gut health and as signalling molecules, SCFA might associated with an increased in vitro production of the
enter the systemic circulation and directly affect substrate metabolism and major SCFA, as measured by the ability of individuals
function of peripheral tissues faecal inoculum to ferment dietary fibres in vitro.39
■■ SCFA might increase intestinal energy harvesting and promote the development
of obesity, but could also increase energy expenditure and anorexic hormone
SCFA absorption and systemic concentrations
production, as well as improving appetite regulation
■■ Increasing evidence supports a beneficial role for SCFA in adipose tissue,
Colonic SCFA absorption is efficient and rapid with only
skeletal muscle and liver substrate metabolism and function, thereby 5–10% of the SCFA excreted in the faeces.27 SCFA can be
contributing to improved insulin sensitivity absorbed through the apical membrane of colonic epi-
■■ Well-controlled human intervention studies investigating the role of SCFA thelial cells via four putative mechanisms: non-ionic diff­
and differential SCFA availability on gut and systemic metabolic health are usion of protonated SCFA;40 exchange with bicarb­onate in
eagerly awaited a ratio of 1:1;41 via the hydrogen-coupled monocarboxylate
transporter 1 (MCT 1), MCT 2 and MCT 4;42,43 and via
sodium-coupled monocarboxylate transporter 1.43
formate, valerate and caproate are present in substan- SCFA absorbed in the caecum, ascending colon or
tially lower amounts and make up the remaining <5%.21 transverse colon are transported into the superior mesen­
Nondigested proteins or peptides might also be sub- teric vein, whilst SCFA absorbed in the descending colon
strates for microbial production of SCFA.22–24 However, and sigmoid enter the inferior mesenteric vein, which
microbial protein fermentation yields a diverse range both drain into the portal vein and liver.44,45 Interestingly,
of metabolites, including branched-chain fatty acids, SCFA absorbed in the sigmoid and rectal region can also
amines, phenols, indoles and thiols, many of which bypass the liver via the pelvic plexus, which drains into
might have negative consequences on epithelial cell the inferior vena cava, thereby reaching the systemic
metabolism and barrier function and affect the host’s circulation directly.45 In the available human studies,
gut and metabolic health.20,24–26 in which SCFA dynamics have been measured during
Most SCFA production occurs in the caecum and major upper abdominal surgery, acetate is absorbed in
proximal colon.27 The concentrations of faecal acetate, the highest amount in the liver and reaches the highest
propionate and butyrate are found in a molar ratio of peripheral concentrations (19–160 μmol/l) compared with
approximately 3:1:1.28,29 Data from indivi­duals who have propi­onate (1–13 μmol/l) and butyrate (1–12 μmol/l).44,46
died suddenly showed that the SCFA concentrations are Systemic concentrations of SCFA depend on both the
approximately 10-fold higher in the caecum (131 mmol/kg production and absorption rates in the gut, which in turn
of luminal content) than in the ileum (13 mmol/kg). SCFA relate to the dietary pattern. For example, in a 4‑week
concentrations were 123 mmol/kg, 117 mmol/kg, human intervention study, supplementation with resist-
80 mmol/kg and 100 mmol/kg in the luminal content ant starch (30 g per day), increased postprandial systemic
in the ascending, transverse, descending and sigmoid acetate and propionate concentrations up to 280 μmol/l
colon, respectively.21 The SCFA molar ratio was similar and 13 μmol/l, respectively.47
in samples from the caecum, the colonic regions and
faecal samples.21 SCFA receptors
Theoretically, ~400–800 mmol SCFA can be produced Butyrate is the major intestinal fuel, 48 supplying
per day with a high-fibre diet, assuming that 10 g of ~60–70% of the energy needs of isolated colonocytes.49,50
dietary fibre fermentation yields ~100 mmol SCFA.30,31 Moreover, acetate, propionate and butyrate stimulate
These estimates agree with results from studies of in vitro epithelial cell proliferation and differentiation in rats.51
fermentation of ileal effluents from patients with an SCFA might affect gut and host metabolism via the
ileostomy, which produced 120–420 mmol acetate and activation of G‑protein coupled cell surface receptors,
45–148 mmol propionate per day, depending upon the G‑protein coupled receptor 41 (GPR41, also known
type and amount of complex carbohydrates in their diet.32 as free fatty acid receptor 3) and GPR43 (also known as
Fermentation of oat bran, which contains high amounts free fatty acid receptor 2).52,53 GPR41 is primarily acti-
β‑glucan, and kidney beans, containing high amounts of vated by propionate followed by butyrate and acetate,
resistant starch and α‑galactosides, yielded the highest GPR43 is activated by all three SCFA at a similar rate.52,53
production of major SCFA.32 Other factors, including GPR41 and GPR43 proteins and mRNA have been found
the specific species of microbiota, diversity and absolute in human colonic tissue.54,55 Notably, both receptors are
amount of host’s intestinal microbiota and gut transit also expressed in human white adipose tissue,19,52,53 skel-
time of food, also have an important role in the produc- etal muscle and liver,52 which indicates that SCFA might
tion of SCFA.27,33,34 During increased speed of transit also affect substrate and energy metabolism directly in
through the colon, substrates available for microbial fer- peripheral tissues. Furthermore, GPR109a (also known as
mentation are diminished, which might lead to reduced hydroxycarboxylic acid receptor 2), which is expressed in

2 | ADVANCE ONLINE PUBLICATION www.nature.com/nrendo


© 2015 Macmillan Publishers Limited. All rights reserved
REVIEWS

compositions have provided a more direct link between


composition of the gut microbiota, SCFA production and
obesity.73,74 In mice with genet­ically induced obesity, an
Low-grade systemic inflammation
Fat oxidation increased ratio of Firmicutes to Bacteroidetes (the two
Lipid storage dominant microbial phyla that can generate SCFA), was
Glycogen storage associated with increased caecal concen­trations of acetate
Glucose uptake
Insulin resistance and butyrate5 or acetate and propionate,17 respectively,
when compared with lean control mice.
Glucose Data from humans also support a relationship between
Proinflammatory cytokines
and chemokines the gut microbiota, SCFA and markers of obesity.28,29,75–77
An increase in faecal concentrations of SCFA (in particu-
lar propionate) and a decreased ratio of Bacteroidetes to
Circulating insulin Firmicutes has been reported in indivi­duals with obesity,
compared with lean individuals.28,29 However, human
data are inconsistent and can variably show increased,
Lipid buffering capacity similar or decreased Firmicutes to Bacteroidetes
Glucose ratios. 29,75,76 Furthermore, when comparing the gut
microbiota of monozygotic twins and dizygotic twins
who are discordant for obesity, the gut microbiome in
the twin with obesity was associated with a dominant
abundance of Firmicutes, reduced microbial diversity and
enrichment of genes involved in carbohydrate fermen­
Lipid overflow:
circulating TG and FA tation.77 Taken together, these data suggest that gut and
Lipid storage faecal SCFA might be increased in human obesity.
Gluconeogenesis By contrast, the data from several animal studies indi-
Glycogen storage cate that treatment with SCFA can reduce or reverse gains
Lipogenesis
Insulin resistance
in body weight and adiposity.11–14,16,78 In obese mice, oral
administration of sodium butyrate leads to loss of body
Figure 1 | Interorgan crosstalk and obesity-induced insulin
Nature resistance. Under
Reviews | Endocrinology weight, via increased energy expenditure and fat oxi­
conditions of positive energy balance (that is, energy intake is much greater than dation.12 In addition, oral acetate, propi­onate and butyrate
energy expenditure), adipose tissue exceeds its buffering capacity to store all administration to mice fed a high-fat diet reduced body
excess energy in the form of TG, which results in overflow of lipids into the weight and improved insulin sensitivity without changing
circulation. This increased lipid supply to nonadipose tissues such as the liver,
food intake or levels of physical activity.16 Furthermore,
skeletal muscle and pancreas results in ectopic fat storage in these tissues and
the development of insulin resistance. Together with a reduced lipid buffering faecal microbiota transplantation from human twins dis-
capacity, adipose tissue becomes inflamed, which results in an increased cordant for obesity to germ-free mice increased adipos-
production and secretion of proinflammatory cytokines and adipokines, such as ity, reduced caecal SCFA and increased monosaccharide
tumour necrosis factor, IL‑6 and C-C motif chemokine 2 (commonly known as and disaccharide concen­trations after a diet rich in plant
monocyte chemoattractant protein‑1), which might also contribute to the polysaccharides in mice receiving microbiota from obese
development of peripheral insulin resistance and a disturbed glucose donors compared with those receiving the lean co-twin
homeostasis. Abbreviations: FA, fatty acid; TG, triglyceride. microbiota.79 These data suggest that the microbiota from
patients with obesity has a lower capacity to completely
gut epithelial cells, adipocytes and immune cells, has been ferment and breakdown complex carbohydrates than
found to respond to butyrate, but not acetate or propion- microbiota from lean individuals.
ate.56–59 In addition, acetate and propionate can also mod- In humans, the beneficial metabolic effects of faecal
ulate blood pressure via olfactory receptor 51E2, which is transplantation from lean donors to men with obesity have
expressed in kidney and vascular smooth muscle cells.60 been associated with an increased content of butyrate-
producing bacteria.9 Further data also provide evidence
SCFA and control of body weight that dietary fibres, such as resistant starch, inulin, oligo­
SCFA and energy harvesting fructose, polydextrose and galactooligo­saccharides are
The gut microbiota in the colon can enable the host to associated with improved weight control and insulin
gain extra energy, mainly through the production of SCFA sensi­tivity. This effect might be triggered via SCFA action
from otherwise indigestible carbohydrates. Since the on the ability to store energy and respond to energy
1940s, in vitro61–64 and in vivo animal65–68 and human69–72 intake via differential mechanisms, such as production
data have been published, in which the role of SCFA in of anorexic hormones, increased energy expenditure and
adiposity has been studied.61–72 Until the 1990s, charac- improved metabolic function of peripheral tissues such as
terization of the gut microbiota was laborious and time- skeletal muscle and adipose tissue.47,80–89
consuming because bacteriological culture was the only These data suggest that gut or faecal SCFA are increa­sed
technique available to specifically charac­terize its compo- in the obesity-associated gut micro­biome, and might indi-
sition.73 Since 1993, advances in culture-independent tech- cate an increased capacity to extract calories from other-
niques to analyse genetic diversity of complex microbial wise indigestible foods in these individuals. Nevertheless,

NATURE REVIEWS | ENDOCRINOLOGY ADVANCE ONLINE PUBLICATION | 3


© 2015 Macmillan Publishers Limited. All rights reserved
REVIEWS

these data are controversial and the evidence that SCFA secretion.106–108 Such diets can also increase expression
production by fermentation accounts for 5–10% of the of PYY and proglucagon (a precursor of glucagon and
daily human energy intake is only circumstantial.90 Taken several other hormones including GLP‑1) in the caecum
together, the importance of energy harvesting from and colon,108,109 and promote L‑cell differentiation in the
indigestible carbohydrates is largely unknown. proximal colon.110
These findings, and the potent anorectic influence of gut
SCFA as a regulator of energy intake hormones, led to an increased number of human studies on
The results from dietary intervention studies in humans the effects of fermentable polysacch­arides and the release of
indicate that SCFA might be used to regulate energy intake PYY and GLP‑1 in relation to control of body weight. For
and body weight. In an acute crossover trial with healthy example, in one study, supplementation with oligo­fructose
volunteers, an intake of 10 g of inulin propionate ester (16 g per day for 2 weeks) lowered subjective hunger
increased satiety and reduced appetite, as measured via ratings (P <0.01), as measured by visual analogue scales,
visual analogue scales for hunger and satiety, compared which was associated with increased post­prandial plasma
with inulin alone.91 Furthermore, in humans with obesity, levels of PYY and GLP‑1 in healthy adults.84 Moreover,
inulin propionate ester intake (10 g per day) versus inulin oligofructose intake for 12 weeks (21 g per day) led to a
alone (10 g per day) for 24 weeks prevented gains in body moderate but significant weight loss (1.03 kg, P <0.01)
weight (percentage of patients who gained ≥3% of their compared with maltodextrin, which was associated with
baseline body weight: 4% versus 25% for inulin propi­onate an enhanced PYY release and reduced energy intake of
ester versus inulin alone, respectively [P <0.05]; percentage 29% (P <0.01), in patients who were overweight or with
of patients who lost ≥5% of their baseline body weight: 0% obesity (BMI >25 kg/m2).89 Interestingly, an elegant 5‑week
versus 17%, for inulin propionate ester versus inulin alone dose-escalation study in healthy human volunteers high-
respectively [P <0.05]).92 Finally, patients with obesity lighted that intake of at least 35 g per day of oligo­fructose
who consumed acetate in a 500 ml beverage every day for was needed to significantly elevate 8 h post­prandial
12 weeks experienced reduced body weight (74.2 ± 11.0 kg PYY concentrations (17,846 ± 2,537 pmol min/l versus
and 74.6 ± 11.3 kg versus 73.1 ± 8.6 kg and 71.2 ± 8.3 kg, for 22,349 ± 2,527 pmol min/l for 0 g versus 35 g oligofructose
0 g and 1.5 g of acetate, respectively; P <0.05). In this study, respectively; P <0.01).111
total body fat percentage (1.0% versus –3.5%; P <0.05), and Although these data provide evidence that SCFA have a
visceral fat percentage (4.4% versus –4.9%; P <0.05) were role in energy intake via the gut-derived satiety hormones
also reduced following consumption of 1.5 g of acetate.93 PYY and GLP‑1, germ-free mice (which lack microbial
SCFA production) have increased rather than decreased
SCFA control energy intake via gut–brain axis basal plasma levels of GLP‑1 and delayed intestinal transit
One of the mechanisms underlying the effect of SCFA on compared with conventionally raised control mice.112
food intake and satiety might relate to the release of gut- These results suggest that an adaptive mechanism to inade­
derived satiety hormones, in particular glucagon-like quate energy availability in the colon delays gut transit and
peptide‑1 (GLP‑1) and peptide YY (PYY). These proteins enables increased nutrient absorption.112 Consequently,
are secreted by intestinal enteroendocrine L‑cells, which are further human studies on the effects of physiologically rele­
found at the highest density in the ileal and colonic epi- vant SCFA mixtures on the secretion of GLP‑1 and PYY
thelium.94,95 PYY affects appetite and satiety by suppressing and their metabolic consequences should be performed.
neuropeptide Y (NPY) and activating proopiomelano­cortin
(POMC) neurons in the hypo­thalamic arcuate nucleus Acetate and appetite regulation
(ARC), or by delaying gastric emptying.96,97 In addition to its In an elegant mouse study using PET-CT, intravenously
role as an incretin, GLP‑1 also regulates appetite via effects and colonically administered 11C-acetate can cross the
on POMC and NPY neurons in the ARC and is known to blood–brain barrier to be taken up in the hypothalamus.14
inhibit gastric emptying and gastric acid secretion.98–100 This uptake resulted in decreased food intake through
In several in vitro studies using intestinal cell lines from appetite suppression accompanied by increased produc-
rodents and humans, SCFA stimulated the secretion of PYY tion of lactate and γ-aminobutyric acid.14 Furthermore,
and GLP‑1 from L‑cells in a GPR41 and GPR43 depend- in humans, acetate can also cross the blood–brain barrier
ent manner.92,101–105 Additionally, the use of rodent knock- and can also be metabolized in the brain.113 Enhancing
out models has highlighted the importance of the SCFA circulating concentrations of acetate using specific dietary
receptor GPR43 in GLP‑1 and PYY secretion. Knock out fibres could, therefore, affect central appetite regulation
of GPR43 in mice lowers in vivo basal levels of active GLP‑1 and subsequent control of body weight. However, further
by ~43% (P <0.01) and blunts GLP‑1 levels after glucose research is required to investigate whether acetate elicits
gavage by ~47% (P <0.01), compared with wild-type litter­ a central hypothalamic effect in humans.
mates.101 GPR43 deficiency in mice also reduces portal
vein levels of PYY and GLP‑1 following intra-colonic SCFA control energy intake via leptin
propi­onate administration (180 mmol/l) compared with SCFA might stimulate the secretion of the satiety hormone
wild-type littermates.102 leptin (which is derived from adipose tissue), as shown
Furthermore, in in vivo animal studies, diets contain- in mice114 and bovine adipocytes in vitro.115 Additionally,
ing fermentable carbohydrates, (such as oligofructose, 24 h incubation of human subcu­taneous and omental
inulin and resistant starch) can increase PYY and GLP‑1 adipocytes with high physiological concentrations of

4 | ADVANCE ONLINE PUBLICATION www.nature.com/nrendo


© 2015 Macmillan Publishers Limited. All rights reserved
REVIEWS

propionate (3 mmol/l) can stimulate expression of leptin humans is circumstantial. However, observational and
mRNA and the hormone’s secretion.19 Furthermore, oral intervention studies both provide evidence for beneficial
or intravenous administration of propionate can increase effects of SCFA on control of body weight by increasing
expression of leptin mRNA by 45% (P <0.05) in sheep sub­ energy expenditure and stimulating the secretion of ano-
cutaneous adipose tissue.116 Consequently, in addition to rexic hormones. In addition, acetate might directly affect
PYY and GLP‑1, and the direct effects of acetate on the satiety via a central mechanism in the central nervous
central nervous system, the secretion of leptin might partly system; whereas propionate might influence control of
explain the putative satiety-inducing effects of SCFA. body weight via sympathetic nervous system activity and
Nevertheless, usually obesity is characterized by both induction of intestinal gluconeogenesis (as summarized in
increased leptin concentrations and by reduced action of Figure 2). Long-term human dietary intervention studies
the hormone (that is, leptin resistance).117–119 Whether an are warranted to examine the chronic effect of SCFA
SCFA induced increase in leptin secretion might have the supplementation on control of body weight.
ability to overcome leptin resistance, and thereby affect
satiety, remains to be determined. SCFA and adipose tissue function
Lipolysis
SCFA as a regulator of energy use In the late 1960s, the putative antilipolytic effect of acetate
SCFA might also beneficially affect the control of body was recognized in a cross-over study in humans, in which
weight by influencing energy expenditure. In obese orally administered sodium acetate significantly decreased
mice, oral administration of sodium butyrate leads to plasma concentrations of free fatty acids (FFA) by 25%
loss of body weight, via an increased energy expenditure (P <0.01).69 Two decades later, acute rectal admini­stration
and fat oxidation.12 This effect might be partly related of acetate (180 mmol/l) plus propionate (60 mmol/l)
to the increased gene and protein expression of two resulted in a 40% fall in serum levels of FFA (P <0.05) in
thermogenesis-­related genes, peroxisome prolif­erator- healthy humans.72 The latter finding, and the obser­vation
activated receptor γ coactivator 1 α and uncoup­ling that in humans SCFA receptors (GPR41 and GPR43)
protein 1, in brown adipose tissue.12 In addition, oral are expressed at higher levels in subcu­taneous than in
acetate administration to mice fed a high-fat diet can omental adipose tissue, has led to renewed interest in the
decrease total body fat content and hepatic fat accumu- lipolytic properties of SCFA.19,52 In a 2008 study, treatment
lation without changing food intake,120 which has been of differ­entiated murine 3T3‑L1 adipocytes with acetate
linked to an increased expression of thermogenesis-related and propionate (0.1–0.3 mmol/l) reduced intracellular
proteins (peroxisomal acyl-coenzyme A oxidase 1 [also lipolytic activity by up to 50% via activation of GPR43.123
known as AOX], carnitine O-palimtoyltransferase 1, liver A decreased hormone-sensitive lipase phosphorylation at
isoform and mitochondrial uncoupling protein 2) in the Ser563 might underlie this antilipolytic effect, as shown in
liver.120 Whether these effects also translate into the human mature 3T3‑L1 adipocytes treated with supraphysiological
condition remains to be determined. concentrations of acetate (4 mmol/l).124
Furthermore, SCFA potentially regulate body weight by By contrast, incubation of 3T3‑L1 adipocytes with
modulating sympathetic activity and intestinal gluconeo­ supra­physiological concentrations of butyrate (5 mmol/l)
genesis. Administration of propionate can increase energy and propionate (20 mmol/l) resulted in enhanced gly­
expenditure and heart rate in wild-type mice, whereas cerol release, which is indicative of an increased lipo­
these effects were blunted when a β‑adrenergic receptor lytic response.125 These results indicate a potent role of
blocker was administered.121 Additionally, propionate SCFA on lipolytic activity in murine 3T3‑L1 adipo­cytes.
can induce the secretion of norepinephrine from sympa- However, whether these findings extend to human adipo­
thetic neurons, which indicates that propionate increases cytes, in which both GPR41 and GPR43 are expressed (in
sympathetic activity at the ganglion level.121 contrast to 3T3‑L1 adipocytes that only express GPR43)
In an interesting study using mice, the beneficial meta- remains to be determined.126–130
bolic effects of a diet enriched in propionate and butyrate, In addition to intracellular (that is, endogenous) lipo­
including improved glucose tolerance, insulin sensi­tivity lysis, extracellular lipolysis mediated by lipoprotein lipase
and body weight, were completely abolished in mice defi- (LPL) might also be affected by SCFA in adipose tissue.
cient in intestinal gluconeogenesis.11 The investi­gators sug- Incubation of omental adipose tissue explants from
gested that intestinal gluconeogenesis induction promotes overweight humans with propionate (3 mmol/l for 24 h)
release of glucose in the portal vein (acting as an agonist resulted in an increased expression of LPL mRNA.18,131
for low affinity sodium-glucose cotransporter [also known Consistent with this observation, short-term (30 min)
as sodium-glucose co-transporter 3]), which resulted in intravenous infusions of 1.2 mol/l propionate at a rate
decreased hepatic glucose production and increased of 64 μmol/min per kg body weight in sheep increased
satiety and energy expenditure through a brain-related adipose tissue LPL expression by approximately five-
mechanism.11,122 Human in vivo intervention studies are, fold compared with saline (P <0.05).116 Interestingly,
therefore, warranted to examine if these animal data can expression and activation of angio­poietin-like protein 4
be translated to humans. (encoded by ANGPTL4), which inhibits LPL activity, were
In summary, gut microbiota have been implicated in also enhanced in intestinal and hepatic cancer cell lines
energy harvesting and the development of obesity. The when treated with propionate and butyrate, but not when
currently available data are controversial and evidence in treated with acetate.132 However, the actual effect of SCFA

NATURE REVIEWS | ENDOCRINOLOGY ADVANCE ONLINE PUBLICATION | 5


© 2015 Macmillan Publishers Limited. All rights reserved
REVIEWS

Lipolysis AMPK activity


Adipogenesis Insulin sensitivity
Inflammation Lipid accumulation
Leptin

Inflammation
FA

AMPK activity
Insulin sensitivity Satiety
Lipid accumulation
Acetate
Propionate
Butyrate

Intestinal
glucogenesis PYY
Sympathetic activity GPR41/43 GLP-1

Acetate, propionate, butyrate Faeces

Undigested foods

Figure 2 | SCFA and interorgan crosstalk. Fermentation of indigestible foods in the distal intestine results in the production
Nature
of SCFA. The ratio of acetate to propionate to butyrate in the ileum, caecum and colon is ~3:1:1. Reviews
Butyrate and| Endocrinology
propionate
are generally metabolized in the colon and liver and, therefore, mainly affect local gut and liver function. In the distal gut,
SCFA bind to GPR41 and GPR43, which leads to the production of the gut hormones PYY and GLP‑1 and affects satiety and
glucose homeostasis. Furthermore, propionate and butyrate might induce intestinal gluconeogenesis and sympathetic
activity, thereby improving glucose and energy homeostasis. Small amounts of propionate and butyrate and high amounts
of acetate reach the circulation and can also directly affect peripheral adipose tissue, liver and muscle substrate
metabolism and function. In addition, circulating acetate might be taken up by the brain and regulate satiety via a central
homeostatic mechanism. Whether metabolic effects are mainly explained by direct effects of SCFA or indirectly via gut-
derived signalling molecules still remain unclear. Solid lines indicate direct SCFA effects and dashed lines indicate indirect
SCFA effects. Abbreviations: AMPK, adenosine monophosphate-activated protein kinase; FA, fatty acid; GLP‑1, glucagon-like
peptide‑1; GPR, G‑protein coupled receptor; PYY, peptide YY; SCFA, short-chain fatty acid.

on adipocyte and adipose tissue ANGPTL4 expression and abolished when GPR43 expression was blocked using
activity is currently unknown. siRNA,126 which indicates that GPR43 is a major medi­
In summary, acetate and propionate might both inhibit ator of the SCFA adipogenic potential. In addition, in mice
intracellular lipolysis, while propionate can also increase fed a high-fat diet, adipo­genesis was increased in subcu­
the lipid buffering capacity of adipose tissue by increasing taneous adipose tissue, which was associated with an over­
LPL-mediated triglyceride extraction. This effect results expression of GPR43 and PPARγ target genes, namely those
in reduced lipid overflow and decreased ectopic accumu­ that encode CD36 and LPL mRNA.135 By contrast, in vitro
lation of fat, thereby positively affecting insulin sensitivity. differ­entiation of human omental preadipocytes with
However, the role of SCFA on LPL or angiopoietin-like acetate and propionate did not modulate the expression
protein 4 signalling and the impact of SCFA on lipo­ of PPARγ target genes, namely those that encode GPR43
lysis in human adipose tissue cells needs to be investigated and FABP4 (commonly known as aP2), which are known
in detail. to be late markers of adipocyte differentiation, suggesting
that adipocyte differentiation in humans is not mediated
Adipogenesis by GPR43.136
In vitro treatment of murine and porcine preadipocytes These data derived from rodent models indicate that
with sodium acetate, butyrate and propionate enhances acetate, propionate and butyrate might be involved in adipo­
adipo­c yte differentiation, which suggests that SCFA genic differentiation. However, further mechan­istic research
have proadipo­genic properties.133,134 In addition, 3T3‑L1 is required to confirm these outcomes in humans.
preadipo­cytes treated with acetate (0.1 μmol/l) and propi­
onate (0.1 μmol/l) for 7 days enhanced adipo­cyte differ- Adipose tissue inflammation
entiation as measured by oil red O staining, via increased Systemic inflammation, characterized by increased
expression of GPR43 and peroxisome proliferator-activated numbers and tissue infiltration of immune cells, is strongly
receptor γ (PPARγ), a transcription factor involved in early associated with insulin resistance and T2DM. 137,138
adipogenic differentiation.126 These effects were completely Although the stimulus for persistent immune activation

6 | ADVANCE ONLINE PUBLICATION www.nature.com/nrendo


© 2015 Macmillan Publishers Limited. All rights reserved
REVIEWS

remains unclear, adipose tissue is increasingly recognized TREG cells, which was dependent on GPR43.143 Consistent
as a major source of proinflammatory adipocytokines that with this finding, butyrate induced the differentiation
affect cardiometabolic function.1,139 To what extent host of colonic TREG cells both in vivo and in vitro in mice.144
metabolism is controlled by the intestinal immune system Furthermore, butyrate and propionate in drinking water
in close interaction with our gut microbiota and micro- (but not acetate) accelerated extrathymic peripheral dif-
bial products is currently unknown. However, several ferentiation of anti-inflammatory TREG cells in mice treated
SCFA-dependent mechanisms have been proposed. with antibiotics.144 In vitro incubation of naive CD4+ TREG
Regulatory T cells (TREG cells) are central regulators of cells and dendritic cells with butyrate support these out-
antimicrobial immunity and tissue inflammation, and comes,145 which suggests that SCFA have a direct role in
their numbers are decreased in visceral adipose tissue from the regulation of T cells in both the gut and peripheral
people with obesity.140 Disturbances in balance between tissues. However, the role of SCFA in humans on sys-
proinflammatory type 1 T helper cells (TH1, CD4+) cells temic and adipose tissue T‑cell number and activity is
and cytotoxic T cells (CD8+) versus anti-inflammatory currently unknown.
TREG cells are thought to be central in obesity-associated SCFA might also enhance intestinal barrier function,
macrophage recruitment and adipose tissue and systemic giving further support for their anti-inflammatory poten-
inflammation.141,142 Interestingly, treatment of mice with tial. In several studies using intestinal cell lines, SCFA
all three SCFA (acetate, propionate and butyrate) alone (particularly butyrate), improve epithelial barrier func-
or in combination for 3 weeks augmented the number tion and gut permeability, by modulating expression of
and suppressive function of colonic anti-inflammatory tight junction protein and mucins.146–149 Moreover, in a
mouse study, a diet high in dietary fibres, including guar
gum and increased levels of cellulose, induced the forma-
tion of the NLRP3 inflammasome and led to the cleav-
age of proIL‑18 into IL‑18 in colonic epithelial cells, via
the activation of SCFA receptor GPR43 and the butyrate
receptor GPR109a, thereby preventing disruption in epi-
thelial integrity and dysbiosis.150 An improved gut barrier
function is important to prevent leakage of toxic com-
Lipid storage capacity
pounds produced by pathogenic bacteria into the circu­
lation. Metabolic endotoxemia, particularly an increase in
PPARγ
circulating lipopolysaccharide, is associ­ated with chronic
TG P low-grade inflammation, adipose inflammation and
HSL HSL
dysfunction, weight gain and insulin resistance.151–155
TG
Adipogenesis SCFA might also directly counteract lipopolysaccharide-
Glycerol + FFA
induced inflammation. In vitro, acetate and butyrate
LPL
Proinflammatory can decrease lipopolysaccharide-stimulated release of
cytokines and
chemokines
tumour necrosis factor (TNF) from human neutrophils
by ~33% and ~75%, respectively (P <0.01),156 and inhibit
GPR43 lipopolysaccharide-induced nuclear factor κB (NF-κB)
in lipopolysaccharide-stimulated murine macro­phage cell
ANGPTL4
lines.157 In addition, in another in vitro study using human
Propionate Propionate Propionate Propionate peripheral blood mononuclear cells, overnight incubation
Acetate Acetate Acetate with acetate, propionate and butyrate in concentrations
Butyrate Butyrate
of 0.2–100 mmol/l inhibited lipopolysaccharide-induced
production of TNF and IFN‑γ.158 Moreover, in vivo data
showed that acute rectal and intravenous admini­stration
Gastrointestinal tract of sodium acetate (compared with saline) decreased
plasma levels of TNF in women with obesity.15
Nature
Figure 3 | The effect of SCFA on adipose tissue function. Propionate Endocrinology
Reviews |might
Few studies have been conducted to assess the
increase FFA uptake, possibly by affecting the LPL inhibitor ANGPTL4. Acetate
and propionate might also attenuate intracellular lipolysis via decreased HSL
direct effects of SCFA on adipose tissue inflammation.
phosphorylation, via GPR43. In addition, acetate, propionate and butyrate might Propionate treatment (3 mmol/l) reduced levels of the
increase PPARγ-mediated adipogenesis, putatively regulated by a GPR43-related mRNA, but not the protein expression, of the proinflam-
mechanism. Together, these effects might contribute to increased storage of TG matory cytokine resistin in human omental and subcu­
in adipose tissue and a reduced systemic FFA release. Acetate, and especially taneous adipose tissue cells.19 In addition, 24 h incubation
propionate and butyrate, might reduce the secretion of proinflammatory cytokines of human adipose tissue explants with 3 mmol/l propi­onate
and chemokines, thereby possibly reducing local macrophage infiltration. The solid resulted in a reduced mRNA expression and secretion of
lines indicate this evidence is consistent between in vitro and in vivo data. The
the proinflammatory cytokines IL‑4, IL‑10 and TNF, as
dashed lines indicate inconsistent or hypothetical evidence. Abbreviations:
ANGPTL4, angiopoietin-like 4; FFA, free fatty acid; GPR43, G‑protein coupled
well as several chemokines.18 Finally, co-incubation of
receptor 43; HSL, hormone sensitive lipase; LPL, lipoprotein lipase; P, phosphate; murine 3T3‑L1 adipocytes with RAW264.7 macro­phages
PPARγ, peroxisome proliferator-activated receptor γ; SCFA, short-chain fatty acid; and butyrate (0.2–1 mmol/l) for 24 h, reduced TNF, MCP1
TG, triglyceride. and IL‑6 concentrations dose-dependently.159

NATURE REVIEWS | ENDOCRINOLOGY ADVANCE ONLINE PUBLICATION | 7


© 2015 Macmillan Publishers Limited. All rights reserved
REVIEWS

lipo­toxicity and an improvement in liver and skeletal


muscle insulin sensitivity. In addition to an increased
lipid supply, the inability to adjust lipid oxidation to
match supply might contribute to accumulation of bio-
active lipid metab­olites in the muscle, which is associated
PPARδ
with insulin resistance.160,161
Glycogen Only a few studies have examined the direct effects of
repletion FA oxidation
SCFA on muscle fat oxidation, glucose uptake and insulin
pAMPK sensitivity. In obese rats, 6 months of acetate injections
FA (5.2 mg/kg of body weight) improved expression of genes
Glycolysis Lipid storage Lipid buffering
capacity
involved in oxidative and glucose metab­olism (such as,
Mb [which encodes myoglobin] and the glucose trans-
Glucose uptake Insulin action
porter Glut4) and increased muscle adeno­sine mono-
phosphate-activated protein kinase (AMPK) activity in
GPR41/43 GLUT4
skeletal muscle.162 Dietary acetate supple­mentation for
6 days in rats increased glycogen storage and decreased
Acetate GLP-1 Acetate glycolysis in the gastro­cnemius muscle compared with
Butyrate PYY Propionate chow-fed rats.163 In diet-induced obese C57BL/6J mice,
16 weeks of a sodium butyrate enriched high-fat diet
increased the proportion of type 1 oxidative muscle fibres
GPR41/43 and PPARδ expression, which resulted in enhanced mito-
chondrial fat oxidation compared with control mice on
Gastrointestinal tract SCFA a high-fat diet.12 Taken together, these data suggest that
Figure 4 | SCFA and skeletal muscle substrate metabolism. Acetate and acetate and butyrate can improve skeletal muscle glucose
Nature
propionate might reduce ectopic lipid storage in skeletal muscle, due| Endocrinology
Reviews to a and oxidative metab­olism and might increase muscle
decreased lipid supply. Acetate and butyrate increase muscle FA oxidation, possibly lipid turnover, which in turn will improve insulin sensi-
mediated via increased activation of AMPK to pAMPK and a PPARδ-dependent
tivity. Part of the SCFA-induced effects on skeletal muscle
mechanism. In addition, acetate and butyrate might influence skeletal muscle
glucose metabolism in an AMPK-dependent manner, which might increase glucose substrate oxidation might be mediated via SCFA receptor
uptake (via GLUT4) and glycogen storage possibly via a GPR41/GPR43-mediated signalling, as GPR41 and GPR43 are both expressed in
mechanism. SCFA might also indirectly affect muscle insulin sensitivity and human skeletal muscle.52,164 Future investi­gations should
glucose metabolism via increased systemic levels of gut-derived PYY and GLP‑1, also focus on skeletal muscle SCFA uptake and their con-
thereby affecting skeletal muscle insulin action and glucose uptake, contributing to tribution to muscle oxidative metabolism in humans, in
improved muscle insulin sensitivity and glucose handling. The solid lines indicate addition to the role of these receptors.
that evidence is consistent between in vitro and in vivo data. The dashed lines
SCFA might also indirectly affect muscle insulin sensi-
indicate inconsistent evidence or hypothetical mechanisms. Abbreviations: AMPK,
adenosine monophosphate-activated protein kinase; FA, fatty acid; GLP‑1,
tivity and glucose metabolism through gut-derived GLP‑1
glucagon-like peptide‑1; GLUT4, glucose transporter type 4; GPR, G‑protein coupled secretion and modulate muscle microvascular blood
receptor; pAMPK, phosphorylated AMPK; PPARδ, peroxisome proliferator-activated volume and flow, which is associated with augmented
receptor δ; PYY, peptide YY; SCFA, short-chain fatty acid. muscle insulin action and increased muscle glucose utili­
zation.165 SCFA-stimulated secretion of gut-derived PYY
In summary, the currently available data indicate might also improve insulin-mediated glucose uptake in
that SCFA might increase the lipid buffering capa­ skeletal muscle and increase whole-body fat oxidation,
city of adipose tissue by inhibiting intracellular lipo­ as shown in animal studies.166–168 Manipulating intestinal
lysis, increasing LPL-mediated triglyceride extraction SCFA concentrations and ratios might, therefore, bene­
and increasing adipogenic differentiation. Acetate, ficially affect glucose homeo­stasis by triggering the release
propionate and butyrate might all prevent chronic low- of gut-derived incretins and increasing skeletal muscle
grade inflammation by upregulating anti-inflammatory insulin action (Figure 4). To date, no human intervention
TREG cells, decreasing metabolic endotoxemia and redu­ study on the direct associ­ation between gut-derived SCFA
cing production of proinflammatory adipocytokines and skeletal muscle function has been performed.
and chemokines. These effects will influence the recruit-
ment and accumulation of monocytes in adipose tissue SCFA and liver function
(Figure 3). Currently, most of the evidence for these Obesity-induced hepatic fat accumulation and chronic
functions is derived from in vitro studies using animal- low-grade inflammation are strongly associated with
derived cell models, which do not directly recapitulate the insulin resistance. Consequently, factors that improve
situation in human diseases. Consequently, future investi­ hepatic substrate metabolism might be used to prevent
gations should focus on human SCFA metabolism using or treat insulin resistance and T2DM.
human in vivo and cell models. In a study in which radiolabelled 13C-SCFA was
directly infused in the caecum of mice, acetate and
SCFA and skeletal muscle function butyrate were found to be highly involved in liver pal-
The decline in circulating lipid concentrations might mitate and chole­sterol synthesis.169 However, a large
contribute to a reduction in ectopic fat storage, part of the infused propionate was used for de novo

8 | ADVANCE ONLINE PUBLICATION www.nature.com/nrendo


© 2015 Macmillan Publishers Limited. All rights reserved
REVIEWS

gluconeogenesis, and not lipogenesis.169 Consistent with bacteria can prevent the progression of diet-induced
this observation, in isolated rat hepatocytes, acetate acts non­a lcoholic fatty liver disease and improve insulin
as a lipogenic substrate whilst propionate suppresses resistance and triglyceride content, which is possibly
lipogenesis through decreased expression of fatty mediated by increased activation of liver AMPK.176 The
acid synthase.170 effects of SCFA on liver glucose and lipid metab­olism
Additionally, SCFA might affect hepatic glucose and might be mediated in part by GPR41 and GPR43, which
lipid metabolism through an AMPK-dependent mecha- are expressed in human hepatic cells.52,164
nism.171,172 Bovine hepatocytes, cultured with acetate in In summary, propionate might be a direct substrate for
concentrations of 1.8–7.2 mmol/l for 3 h increased the liver gluconeogenesis and acetate and butyrate for lipo­
AMP to ATP ratio and subsequent AMPK phosphory­ genesis. Furthermore, acetate and butyrate in particular
lation, increasing the expression of PPARα target genes might indirectly affect liver function and metabolism
involved in lipid oxidation.173 Moreover, oral and intra­ in an AMPK-dependent manner (Figure 5). However,
venous administration of acetate,16,120,174,175 butyrate16 studies in humans that investigate the influence of
and propionate16 in animal models of obesity and T2DM SCFA effects on parameters of liver metabolism are
can decrease the accumulation of lipids in the liver and currently lacking.
improve glucose tolerance. This mechanism functioned
via increased hepatic AMPK phosphorylation and expres- SCFA and glucose homeostasis in humans
sion of PPARα target genes involved in FFA oxidation, The putative direct and indirect effects of SCFA on the
glycogen storage, thermogenesis, gluconeo­genesis and function of peripheral tissues (Figure 2) suggest that
lipogenesis. Taken together, these data suggest a poten- SCFA might be strongly involved in glucose regulation
tial role for SCFA in hepatic lipid and glucose handling. in humans. Several studies have focused on the acute
Moreover, treatment of rats with butyrate-producing effects of SCFA administration on parameters of glucose
and insulin metabolism (Table 1). Blood glucose and
insulin levels are unaffected after rectal admini­stration of
acetate and propionate within 30 min, but in this investi­
gation, both SCFA decreased circulating concentrations
of FFA.72 However, in a subsequent study, acute rectal
acetate infusions increased circulating levels of gluca-
gon and decreased circulating levels of FFA, but did not
affect blood glucose or insulin levels.177 Rectal propi­onate
infusions increased glucose and glucagon levels, but did
PPARα not affect FFA or insulin concentrations, which suggests
Glycogen that gut-derived propionate might be a substrate for
storage FA oxidation
gluconeogenesis (Table 1).177
AMP/ATP pAMPK
However, other investigators have found no effects on
glucose metabolism after 3 h gastric infusions of acetate
FA
Gluconeogenesis
Lipid storage Lipid buffering and/or propionate, although these infusions reduced
capacity
circulating concentrations of FFA.178 No effects of acute
FAS Lipogenesis intra­venously and rectally infused sodium acetate have
been observed on glucose and insulin levels, whilst
intravenous and rectal acetate infusions lowered plasma
GPR41/43
levels of TNF and rectal acetate infusions increased
Acetate Propionate Acetate Acetate PYY levels.15 Finally, no differences have been seen in
Propionate
Butyrate Butyrate the clearance of intra­venously-administered acetate in
adults who were healthy or had hyperinsulinaemia.179
However, in both these groups of individuals, levels of
FFA fall after acetate administration and a greater FFA
Gastrointestinal tract fall and rebound was observed in the healthy adults
Figure 5 | SCFA and liver function. SCFA might act asNature for | Endocrinology
Reviews
substrates compared with those with hyperinsulinaemia.179 A fall
gluconeogenesis and de novo lipogenesis. Propionate acts in the liver as a and rebound of FFA were both negatively correlated
precursor for de novo gluconeogenesis and attenuates lipogenesis via inhibition with insulin resistance indices (such as HOMA and the
of FAS expression. Acetate and butyrate are involved in liver lipogenesis. Moreover, Insulinogenic Index), which indicates an association
acetate and butyrate might directly increase hepatic AMPK phosphorylation and between an acetate-induced decrease in lipolysis and
activity, via an increase in the ratio of AMP to ATP, and the upregulation of PPARα improved insulin resistance (Table 1).179
target genes, thereby increasing FA oxidation and glycogen storage, which is
Data on the long-term administration of SCFA are
possibly mediated via GPR41/GPR43-dependent mechanism. The solid lines
indicate this evidence is consistent between in vitro and in vivo data. The dashed
limited (Table 1). In one study, propionate supple-
lines indicate inconsistent or hypothetical evidence. Abbreviations: AMPK, mentation for 7 weeks led to decreased fasting glucose
adenosine monophosphate-activated protein kinase; FA, fatty acid; FAS, fatty acid levels and maximum insulin increments during an oral
synthase; GPR, G‑protein coupled receptor; pAMPK, phosphorylated AMPK; PPARα, glucose tolerance test in healthy women.180 Daily propi­
peroxisome proliferator-activated receptor α; SCFA, short-chain fatty acid. onate supple­mentation in bread for 1 week decreases

NATURE REVIEWS | ENDOCRINOLOGY ADVANCE ONLINE PUBLICATION | 9


© 2015 Macmillan Publishers Limited. All rights reserved
REVIEWS

Table 1 | SCFA intervention studies in humans


Participants SCFA Design Effects on glucose homeostasis Study
9 healthy adults and Intravenous sodium acetate Acute comparison of Acetate-induced FFA response was ~14% Fernandes et al.
9 adults with (140 mmol/l in 90 min) healthy versus and FFA rebound ~30% greater in healthy (2012)179
hyperinsulinaemia hyperinsulinaemia adults compared with patients who were
hyperinsulinaemic (P <0.01)
6 women who were Sodium acetate Acute; placebo controlled No significant changes in glucose and Freeland et al.
overweight or obese 60 mmol in 300 ml water; (saline) insulin levels (2010)15
(BMI >25 kg/m2) with 200 mmol/l (rectally), Increase in PYY and decrease in TNF
hyperinsulinaemia Intravenous 20 mmol in 100 ml
water in 5 min
6 healthy adults Acetate (12 mmol/h), Acute; placebo controlled No changes in blood glucose, plasma Laurent et al.
or propionate (4 mmol/h), (saline) insulin concentrations or hepatic glucose (1995)178
or acetate + propionate production
(12 mmol/h + 4 mmol/h) Decreased circulating FFA after all SCFA
for 3 h (gastric) infusions
6 healthy adults Oral sodium propionate Chronic; placebo 38% lowered 2 h glucose AUC values Todesco et al.
(9.9 g per day) in white bread controlled (standard (P <0.05) and a lower glucose peak (1991)71
for 1 week white bread)
6 healthy adults 800 ml rectal infusions with Acute; placebo controlled Acetate alone increased serum levels Wolever et al.
180 mmol acetate alone, (saline) of glucagon by ~26% (P <0.05) and (1991)177
180 mmol propionate alone or a decreased FFA levels by ~22% (P <0.05),
combination of both (180 mmol but did not affect glucose or insulin levels
acetate + 60 mmol propionate) Propionate alone increased glucose by
~8% (P <0.05), and glucagon by ~13%,
but no effects on insulin levels
were observed
20 healthy female adults Oral sodium propionate Chronic; placebo Reduced fasting blood glucose and Venter et al.
(7.5 g capsules daily) for controlled insulin response (P <0.05) (1990)180
7 weeks (calcium phosphate)
6 healthy adults 800 ml rectal infusions with Acute; placebo controlled No effects on circulating glucose Wolever et al.
90 mmol acetate + 30 mmol (saline) or insulin (1989)72
propionate (90 mmol/l, isotonic) 40% decreased serum levels of FFA after
or 180 mmol acetate + 60 mmol 180 mmol/l SCFA infusion (P <0.05)
propionate (180 mmol/l,
hypertonic) within 30 min
Abbreviations: AUC, area under the curve; FFA, free fatty acid; PYY, peptide YY; SCFA, short-chain fatty acid; TNF, tumour necrosis factor.

2-h postprandial glucose levels after consumption supplemen­tation of a galactooligo­saccharide mixture


of the bread in the morning on day 8, compared with in patients who were overweight increased faecal levels
consumption of a propionate-free bread.71 However, in of Bifidobacteria and decreased fasting concentrations
this study, the glucose-lowering effects were related to of insulin and tri­glyceride.86 In another elegant study,
decreased digestion of the bread-derived starch, as a investi­gators found an improved postprandial glucose
higher faecal bulk was found in the propionate-treated response, and a shift towards a butyrate-producing gut
group than in the control group.71 Further indications microbiota composition after 3 months of intake of a
for a beneficial effect of SCFA on insulin sensitivity and mixture of inulin and oligo­f ructose (16 g per day) in
glucose homeostasis are derived from human dietary women with obesity.88 However, no effects were found on
intervention studies using fermentable polysaccharides HbA1c, HOMA and fasting levels of insulin and glucose
(Table 2). A daily dose of 30 g resistant starch supple- in this study.88
mentation for 4 weeks can enhance whole-body and These data provide direct and indirect evidence for a
skeletal muscle insulin sensitivity in healthy individuals, role of SCFA in glucose control and insulin sensitizing.
which was accompanied by increased systemic acetate However, well-controlled long-term studies using fer-
and propionate levels and an increase in adipose tissue mentable dietary fibres have failed to demonstrate any
and skeletal muscle acetate uptake.47 A 2-week treatment beneficial effects on glucose homeostasis.87,111 These con-
with oligofructose (16 g per day) decreases post­prandial flicting outcomes indicate that an urgent need exists for
glucose levels, which was associated with increased human SCFA intervention studies to properly investigate
plasma levels of GLP‑1 and an increased hydrogen the effects on glucose homeostasis and insulin sensitivity.
excretion in the breath (increased microbial fermenta-
tion) in healthy adults (Table 2).84 Consistent with these Conclusions
observations, reduced postprandial glucose and insulin The impact of SCFA on energy homeostasis is ambi­
levels and increased PYY levels have been observed after guous. The results of animal studies suggest that SCFA
oligo­fructose (21 g per day) intake for 12 weeks in adults might be involved in an increased energy harvesting, and
who were overweight (Table 2).89 In addition, a 12-week thereby contribute to development of obesity. Conversely,

10 | ADVANCE ONLINE PUBLICATION www.nature.com/nrendo


© 2015 Macmillan Publishers Limited. All rights reserved
REVIEWS

Table 2 | Long-term dietary intervention studies with effects on glucose homeostasis and insulin sensitivity
Participants Nondigestible carbohydrate Design Effects on glucose homeostasis Study
10 healthy adults 30 g resistant starch 4 weeks; placebo- Improved whole-body insulin sensitivity Robertson et al.
(10 g three times per day) controlled (euglycemic–hyperinsulinaemic clamp) (2005)47
(20 g digestible starch) by ~13% (P <0.05)
10 healthy adults 16 g oligofructose 2 weeks; placebo- 2 h postprandial glucose AUC was Cani et al. (2009)84
(8 g twice per day) controlled reduced by ~17% (P <0.05)
(16 g dextrin maltose) Increased plasma levels of GLP‑1 and
PYY and increased breath-hydrogen
excretion
48 adults who were 21 g oligofructose 12 weeks; placebo- Absolute 6 h postprandial plasma Parnell et al.
overweight or obese (7 g three times per day) controlled concentrations of glucose and insulin (2009)89
(BMI >25 kg/m2) (7.89 g isocaloric were reduced by 5% (P <0.01) ~10%
maltodextrin) (P <0.01), respectively
30 women with obesity 16 g inulin/oligofructose mix 3 months; placebo- Reduced post-OGTT glucose response Dewulf et al.
(BMI >30 kg/m2) (8 g twice per day) controlled (7%; P <0.01) (2013)88
(16 g maltodextrin) No effects on HOMA, fasting glucose and
insulin and HbA1c
45 adults who were 5.5 g galactooligosaccharide 12 weeks; placebo- Decreased fasting insulin (~14% Vulevic et al.
overweight or obese mixture (once a day) controlled (P <0.01), triglyceride and C‑reactive (2013)86
(BMI >25 kg/m2) (5.5 g maltodextrin) protein plasma concentrations
12 healthy adults 15 g, 25 g, 35 g, 45–55 g 5 weeks; No effects on fasting or postprandial Pedersen et al.
oligofructose daily (increased dose-escalating pilot glucose and insulin (2013)111
per week)
22 adults who were 30 g oligofructose 6 weeks; 30 g cellulose No effects on fasting or postprandial Daud et al. (2014)87
overweight or obese (10 g three times per day) as control glucose and insulin (neither between
(BMI 25–35 kg/m2) treatments nor within group)
Abbreviations: AUC, area under the curve; GLP‑1, glucagon-like peptide 1; OGTT, oral glucose tolerance test; PYY, peptide YY.

SCFA might increase energy expenditure, stimulate the hepatic substrate metabolism in an AMPK-dependent
production of satiety hormones, and acetate itself might manner. Taken together, these effects might contribute to
enhance central appetite regulation, preventing rather improved hepatic and peripheral insulin sensitivity and
than promoting obesity. Additionally, SCFA, and the glucose homeostasis.
microbiota that produce them, have been associated with The evidence for a potential role of SCFA as a met-
improved insulin sensitivity and metabolic health in, abolic tool to prevent and counteract obesity and
albeit a limited number of, human intervention studies. associ­ated cardiometabolic risk factors such as insulin
Several different mechanisms might be responsible for resistance is increasing. Nevertheless, most data are
the putative positive effects of SCFA on insulin sensi­ derived from animal and in vitro studies and the clini-
tivity. SCFA, especially acetate and propionate, might cal relevance and metabolic impact in humans remain
improve the lipid buffering capacity of adipose tissue to be established. Concluding that modulation of SCFA
via attenuation of intracellular lipolysis and increased production by microbiota via the intake of complex
adipogenesis. SCFA, in particular butyrate, can regu- carbo­hydrates is a useful tool to prevent obesity and
late obesity-induced chronic low-grade inflammation, obesity-related disturbances in glucose metabolism
by activating anti-inflammatory TREG cells and suppress and insulin resistance is probably premature. Long-term
pathways involved in proinflammatory cytokine and human intervention studies that are well-controlled are
chemokine production. In addition, acetate, propi­ needed to clarify the role of SCFA on control of body
onate and butyrate might directly decrease secretion weight and insulin sensitivity.
of adipose tissue-derived proinflammatory cytokines
and chemokines. An improved SCFA-related adipose
Review criteria
tissue function might decrease systemic lipid over-
flow and inflammation, and thereby reduce ectopic fat PubMed was searched for relevant topics, using the
storage in tissues such as the skeletal muscle and liver. following search terms: “short-chain fatty acids”,
“acetate”, “butyrate”, “propionate” or “dietary fiber”
Furthermore, acetate and butyrate might improve local
in combination with “obesity”, “weight”, “satiety”,
muscle fat oxidation in an AMPK-dependent manner “type 2 diabetes”, “glycemic control”, “glucose-lowering
or by changing the oxidative state of muscle fibres, mechanisms”, “energy metabolism”, “FFAR2”, “FFAR3”,
and thereby improve the capacity to utilize and switch “inflammation”, “treg”, “microbiota”, “fermentation”,
between the major fuels lipids and carbo­hydrates (that “intestinal homeostasis”, “cardiovascular disease”
is, enhance metabolic flexi­bility). In the liver, acetate and and “metabolic control”, without time constraints on
butyrate can act directly as a substrate for lipo­genesis, publication. References cited in this article include
whereas propi­onate is mainly a substrate for gluconeo­ primarily English-language original research and some
specific reviews by experts in the field.
genesis. Moreover, acetate and butyrate can regulate

NATURE REVIEWS | ENDOCRINOLOGY ADVANCE ONLINE PUBLICATION | 11


© 2015 Macmillan Publishers Limited. All rights reserved
REVIEWS

1. Shulman, G. I. Ectopic fat in insulin resistance, metabolites on the colonic epithelium and transporter (MCT1) in pig and human colon:
dyslipidemia, and cardiometabolic disease. physiopathological consequences. Amino Acids its potential to transport L‑lactate as well as
N. Engl. J. Med. 371, 1131–1141 (2014). 33, 547–562 (2007). butyrate. J. Physiol. 513, 719–732 (1998).
2. Grundy, S. M. Obesity, metabolic syndrome, 23. Geypens, B. et al. Influence of dietary protein 43. Moschen, I., Bröer, A., Galic, S., Lang, F.
and cardiovascular disease. J. Clin. Endocrinol. supplements on the formation of bacterial & Bröer, S. Significance of short chain fatty acid
Metab. 89, 2595–2600 (2004). metabolites in the colon. Gut 41, 70–76 (1997). transport by members of the monocarboxylate
3. Kahn, S. E., Hull, R. L. & Utzschneider, K. M. 24. Macfarlane, G., Gibson, G., Beatty, E. transporter family (MCT). Neurochem. Res. 37,
Mechanisms linking obesity to insulin resistance & Cummings, J. Estimation of short-chain fatty 2562–2568 (2012).
and type 2 diabetes. Nature 444, 840–846 acid production from protein by human intestinal 44. Bloemen, J. G. et al. Short chain fatty acids
(2006). bacteria based on branched-chain fatty acid exchange across the gut and liver in humans
4. Bäckhed, F. et al. The gut microbiota as an measurements. FEMS Microbiol. Lett. 101, measured at surgery. Clin. Nutr. 28, 657–661
environmental factor that regulates fat storage. 81–88 (1992). (2009).
Proc. Natl Acad. Sci. USA 101, 15718–15723 25. Corpet, D. E. et al. Colonic protein fermentation 45. Araghizadeh, A. & Abdelnaby, A. in Colorectal
(2004). and promotion of colon carcinogenesis by Surgery (eds Bailey, H. R., Billingham, R. P.,
5. Turnbaugh, P. J. et al. An obesity-associated gut thermolyzed casein. Nutr. Cancer 23, 271–281 Stamos, M. J. & Snyder, M. J.) 3–17 (Elsevier
microbiome with increased capacity for energy (1995). Health Sciences, 2012).
harvest. Nature 444, 1027–1031 (2006). 26. Windey, K., De Preter, V. & Verbeke, K. Relevance 46. Bloemen, J. G. et al. Short chain fatty acids
6. Membrez, M. et al. Gut microbiota modulation of protein fermentation to gut health. Mol. Nutr. exchange: Is the cirrhotic, dysfunctional liver still
with norfloxacin and ampicillin enhances glucose Food Res. 56, 184–196 (2012). able to clear them? Clin. Nutr. 29, 365–369
tolerance in mice. FASEB J. 22, 2416–2426 27. Wong, J. M., de Souza, R., Kendall, C. W., (2010).
(2008). Emam, A. & Jenkins, D. J. Colonic health: 47. Robertson, M. D., Bickerton, A. S., Dennis, A. L.,
7. Andersson, U. et al. Probiotics lower plasma fermentation and short chain fatty acids. J. Clin. Vidal, H. & Frayn, K. N. Insulin-sensitizing effects
glucose in the high-fat fed C57BL/6J mouse. Gastroenterol. 40, 235–243 (2006). of dietary resistant starch and effects on
Benef. Microbes 1, 189–196 (2010). 28. Schwiertz, A. et al. Microbiota and SCFA in lean skeletal muscle and adipose tissue metabolism.
8. Vrieze, A. et al. Impact of oral vancomycin on gut and overweight healthy subjects. Obesity 18, Am. J. Clin. Nutr. 82, 559–567 (2005).
microbiota, bile acid metabolism, and insulin 190–195 (2009). 48. Roediger, W. Role of anaerobic bacteria in the
sensitivity. J. Hepatol. 60, 824–831 (2014). 29. Fernandes, J., Su, W., Rahat-Rozenbloom, S., metabolic welfare of the colonic mucosa in man.
9. Vrieze, A. et al. Transfer of intestinal microbiota Wolever, T. & Comelli, E. Adiposity, gut Gut 21, 793–798 (1980).
from lean donors increases insulin sensitivity in microbiota and faecal short chain fatty acids 49. Roediger, W. E. W. in Physiological and Clinical
individuals with metabolic syndrome. are linked in adult humans. Nutr. Diabetes 4, Aspects of Short-Chain Fatty Acids
Gastroenterology 143, 913–916 (2012). e121 (2014). (eds Cummings, J. H., Rombeau, J. L. & Sakata,
10. Topping, D. L. & Clifton, P. M. Short-chain fatty 30. Cummings, J. H. & Macfarlane, G. T. Colonic T.) 337–351 (Cambridge University Press, 1995).
acids and human colonic function: roles of microflora: nutrition and health. Nutrition 13, 50. Ardawi, M. & Newsholme, E. Fuel utilization in
resistant starch and nonstarch polysaccharides. 476–478 (1997). colonocytes of the rat. Biochem. J. 231,
Physiol. Rev. 81, 1031–1064 (2001). 31. McBurney, M. I. & Thompson, L. U. In vitro 713–719 (1985).
11. De Vadder, F. et al. Microbiota-generated fermentabilities of purified fiber supplements. 51. Frankel, W. et al. Mediation of the trophic effects
metabolites promote metabolic benefits via J. Food Sci. 54, 347–350 (1989). of short-chain fatty acids on the rat jejunum and
gut-brain neural circuits. Cell 156, 84–96 (2014). 32. McBurney, M., Thompson, L., Cuff, D. colon. Gastroenterology 106, 375–380 (1994).
12. Gao, Z. et al. Butyrate improves insulin & Jenkins, D. Comparison of ileal effluents, 52. Brown, A. J. et al. The Orphan G protein-coupled
sensitivity and increases energy expenditure dietary fibers, and whole foods in predicting receptors GPR41 and GPR43 are activated by
in mice. Diabetes 58, 1509–1517 (2009). the physiological importance of colonic propionate and other short chain carboxylic
13. Lin, H. V. et al. Butyrate and propionate protect. fermentation. Am. J. Gastroenterol. 83, 536–540 acids. J. Biol. Chem. 278, 11312–11319 (2003).
against diet-induced obesity and regulate gut (1988). 53. Le Poul, E. et al. Functional characterization of
hormones via free fatty acid receptor 33. Venema, K. Microbial metabolites produced human receptors for short chain fatty acids and
3‑independent mechanisms. PLoS ONE 7, by the colonic microbiota as drivers for their role in polymorphonuclear cell activation.
e35240 (2012). immunomodulation in the host. FASEB J. 27, J. Biol. Chem. 278, 25481–25489 (2003).
14. Frost, G. et al. The short-chain fatty acid acetate 643.12 (2013). 54. Karaki, S. et al. Expression of the short-chain
reduces appetite via a central homeostatic 34. Pylkas, A., Juneja, L. & Slavin, J. Comparison of fatty acid receptor, GPR43, in the human colon.
mechanism. Nat. Commun. 5, 3611 (2014). different fibers for in vitro production of short J. Mol. Histol. 39, 135–142 (2008).
15. Freeland, K. R. & Wolever, T. Acute effects of chain fatty acids by intestinal microflora. J. Med. 55. Tazoe, H. et al. Expression of short-chain fatty
intravenous and rectal acetate on glucagon-like Food 8, 113–116 (2005). acid receptor GPR41 in the human colon.
peptide‑1, peptide YY, ghrelin, adiponectin and 35. Lampe, J., Fredstrom, S., Slavin, J. Biomed. Res. 30, 149–156 (2009).
tumour necrosis factor-alpha. Br. J. Nutr. 103, & Potter, J. Sex differences in colonic function: 56. Thangaraju, M. et al. GPR109A is a
460–466 (2010). a randomised trial. Gut 34, 531–536 (1993). G‑protein‑coupled receptor for the bacterial
16. den Besten, G. et al. Short-chain fatty acids 36. Lampe, J., Wetsch, R., Thompson, W. & Slavin, J. fermentation product butyrate and functions as
protect against high-fat diet-induced obesity via Gastrointestinal effects of sugarbeet fiber and a tumor suppressor in colon. Cancer Res. 69,
a PPARγ-dependent switch from lipogenesis to wheat bran in healthy men. Eur. J. Clin. Nutr. 47, 2826–2832 (2009).
fat oxidation. Diabetes 64, 2398–2408 (2015). 543–548 (1993). 57. Liu, F. et al. The expression of GPR109A, NF‑κB
17. Murphy, E. et al. Composition and energy 37. Owens, F. & Isaacson, H. Ruminal microbial and IL‑1β in peripheral blood leukocytes from
harvesting capacity of the gut microbiota: yields: factors influencing synthesis and patients with type 2 diabetes. Ann. Clin. Lab. Sci.
relationship to diet, obesity and time in mouse bypass. Fed. Proc. 36, 198–202 (1977). 44, 443–448 (2014).
models. Gut 59, 1635–1642 (2010). 38. Lewis, S. & Heaton, K. Increasing butyrate 58. Taggart, A. K. et al. (D)‑β-hydroxybutyrate inhibits
18. Al-Lahham, S. H. et al. Propionic acid affects concentration in the distal colon by accelerating adipocyte lipolysis via the nicotinic acid receptor
immune status and metabolism in adipose intestinal transit. Gut 41, 245–251 (1997). PUMA‑G. J. Biol. Chem. 280, 26649–26652
tissue from overweight subjects. Eur. J. Clin. 39. El Oufir, L. et al. Relationships between transit (2005).
Invest. 42, 357–364 (2012). time in man and in vitro fermentation of dietary 59. Tunaru, S. et al. PUMA‑G and HM74 are
19. Al-Lahham, S. H. et al. Regulation of adipokine fiber by fecal bacteria. Eur. J. Clin. Nutr. 54, receptors for nicotinic acid and mediate its anti-
production in human adipose tissue by propionic 603–609 (2000). lipolytic effect. Nat. Med. 9, 352–355 (2003).
acid. Eur. J. Clin. Invest. 40, 401–407 (2010). 40. Ruppin, H., Bar-Meir, S., Soergel, K., Wood, C. 60. Pluznick, J. L. et al. Olfactory receptor responding
20. Macfarlane, G. T. & Macfarlane, S. Bacteria, & Schmitt, M. Jr. Absorption of short-chain fatty to gut microbiota-derived signals plays a role in
colonic fermentation, and gastrointestinal acids by the colon. Gastroenterology 78, renin secretion and blood pressure regulation.
health. J. AOAC Int. 95, 50–60 (2012). 1500–1507 (1980). Proc. Natl Acad. Sci. USA 110, 4410–4415
21. Cummings, J., Pomare, E., Branch, W., Naylor, C. 41. Titus, E. & Ahearn, G. A. Short-chain fatty acid (2013).
& Macfarlane, G. Short chain fatty acids in transport in the intestine of a herbivorous 61. Hellman, B., Larsson, S. & Westman, S. Acetate
human large intestine, portal, hepatic and teleost. J. Exp. Biol. 135, 77–94 (1988). metabolism in isolated epididymal adipose
venous blood. Gut 28, 1221 (1987). 42. Ritzhaupt, A., Wood, I. S., Ellis, A., Hosie, K. B. tissue from obese-hyperglycemic mice of
22. Blachier, F., Mariotti, F., Huneau, J. & Tome, D. & Shirazi-Beechey, S. P. Identification and different ages. Acta Physiol. Scand. 56, 189–198
Effects of amino acid-derived luminal characterization of a monocarboxylate (1962).

12 | ADVANCE ONLINE PUBLICATION www.nature.com/nrendo


© 2015 Macmillan Publishers Limited. All rights reserved
REVIEWS

62. Villee, C. The effect of insulin on the 81. Konings, E., Schoffelen, P. F., Stegen, J. 98. Murphy, K. G. & Bloom, S. R. Gut hormones and
incorporation of C14-labeled pyruvate and & Blaak, E. E. Effect of polydextrose and soluble the regulation of energy homeostasis. Nature
acetate into lipid and protein. Anat. Rec. 101, maize fibre on energy metabolism, metabolic 444, 854–859 (2006).
680 (1948). profile and appetite control in overweight men 99. Näslund, E. et al. GLP‑1 slows solid gastric
63. Feller, D. Metabolism of adipose tissue. and women. Br. J. Nutr. 111, 111–121 (2014). emptying and inhibits insulin, glucagon, and
I. Incorporation of acetate carbon into lipides 82. Reichert, R. G. et al. Decreasing cardiovascular PYY release in humans. Am. J. Physiol. 277,
by slices of adipose tissue. J. Biol. Chem. 206, risk factors in obese individuals using a R910–R916 (1999).
171–180 (1954). combination of PGX®meal replacements and 100. Schjoldager, B., Mortensen, P., Christiansen, J.,
64. Elwood, J., Marcó, A. & Van Bruggen, J. Lipid PGX®granules in a 12-week clinical weight Orskov, C. & Holst, J. GLP‑1 (glucagon-like
metabolism in the diabetic rat. IV. Metabolism of modification program. J. Complement. Integr. peptide 1) and truncated GLP‑1, fragments of
acetate, acetoacetate, butyrate, and mevalonate Med. 10, 135–142 (2013). human proglucagon, inhibit gastric acid
in vitro. J. Biol. Chem. 235, 573–577 (1960). 83. Hashizume, C. et al. Improvement effect of secretion in humans. Dig. Dis. Sci. 34, 703–708
65. Wong, R. K. & Van Bruggen, J. Lipid metabolism resistant maltodextrin in humans with metabolic (1989).
in the diabetic rat. I. Acetate metabolism and syndrome by continuous administration. J. Nutr. 101. Tolhurst, G. et al. Short-chain fatty acids
lipid synthesis in vivo. J. Biol. Chem. 235, 26–29 Sci. Vitaminol. 58, 423–430 (2012). stimulate glucagon-like peptide‑1 secretion via
(1960). 84. Cani, P. D. et al. Gut microbiota fermentation of the G‑protein‑coupled receptor FFAR2. Diabetes
66. Reshef, L., Niv, J. & Shapiro, B. Effect of prebiotics increases satietogenic and incretin gut 61, 364–371 (2012).
propionate on lipogenesis in adipose tissue. peptide production with consequences for 102. Psichas, A. et al. The short chain fatty acid
J. Lipid Res. 8, 682–687 (1967). appetite sensation and glucose response after a propionate stimulates GLP‑1 and PYY secretion
67. Sakata, T. Effects of indigestible dietary bulk and meal. Am. J. Clin. Nutr. 90, 1236–1243 (2009). via free fatty acid receptor 2 in rodents. Int. J.
short chain fatty acids on the tissue weight and 85. Cani, P. D., Joly, E., Horsmans, Y. Obes. 39, 424–429 (2014).
epithelial cell proliferation rate of the digestive & Delzenne, N. M. Oligofructose promotes 103. Reimer, R. A. et al. A human cellular model for
tract in rats. J. Nutr. Sci. Vitaminol. (Tokyo) 32, satiety in healthy human: a pilot study. studying the regulation of glucagon-like peptide‑1
355–362 (1986). Eur. J. Clin. Nutr. 60, 567–572 (2006). secretion. Endocrinology 142, 4522–4528
68. Boillot, J. et al. Effects of dietary propionate on 86. Vulevic, J., Juric, A., Tzortzis, G. & Gibson, G. R. (2001).
hepatic glucose production, whole-body glucose A mixture of trans-galactooligosaccharides 104. Psichas, A. et al. Short chain fatty acids
utilization, carbohydrate and lipid metabolism in reduces markers of metabolic syndrome and stimulate the release of gut hormone peptide YY
normal rats. Br. J. Nutr. 73, 241–251 (1995). modulates the fecal microbiota and immune from human primary enteroendocrine L cells.
69. Crouse, J. R., Gerson, C. D., DeCarli, L. M. function of overweight adults. J. Nutr. 143, Proc. Physiol. Soc. 27, PC331 (2012).
& Lieber, C. S. Role of acetate in the reduction 324–331 (2013). 105. Karaki, S. et al. Short-chain fatty acid receptor,
of plasma free fatty acids produced by ethanol in 87. Daud, N. M. et al. The impact of oligofructose on GPR43, is expressed by enteroendocrine cells
man. J. Lipid Res. 9, 509–512 (1968). stimulation of gut hormones, appetite regulation and mucosal mast cells in rat intestine.
70. Björntorp, P. & Hood, B. Studies on adipose and adiposity. Obesity 22, 1430–1438 (2014). Cell Tissue Res. 324, 353–360 (2006).
tissue from obese patients with or without 88. Dewulf, E. et al. Insight into the prebiotic concept: 106. Cani, P. D., Dewever, C. & Delzenne, N. M. Inulin-
diabetes mellitus. Acta Med. Scand. 179, lessons from an exploratory, double blind type fructans modulate gastrointestinal peptides
221–227 (1966). intervention study with inulin-type fructans in involved in appetite regulation (glucagon-like
71. Todesco, T., Rao, A. V., Bosello, O. & Jenkins, D. obese women. Gut 62, 1112–1121 (2013). peptide‑1 and ghrelin) in rats. Br. J. Nutr. 92,
Propionate lowers blood glucose and alters lipid 89. Parnell, J. A. & Reimer, R. A. Weight loss during 521–526 (2004).
metabolism in healthy subjects. Am. J. Clin. Nutr. oligofructose supplementation is associated with 107. Zhou, J. et al. Dietary resistant starch
54, 860–865 (1991). decreased ghrelin and increased peptide YY in upregulates total GLP‑1 and PYY in a sustained
72. Wolever, T., Brighenti, F., Royall, D., Jenkins, A. overweight and obese adults. Am. J. Clin. Nutr. 89, day-long manner through fermentation in
& Jenkins, D. Effect of rectal infusion of short 1751–1759 (2009). rodents. Am. J. Physiol. Endocrinol. Metab. 295,
chain fatty acids in human subjects. Am. J. 90. Royall, D., Wolever, T. & Jeejeebhoy, K. N. E1160–E1166 (2008).
Gastroenterol. 84, 1027–1033 (1989). Clinical significance of colonic fermentation. 108. Cani, P. D., Neyrinck, A. M., Maton, N.
73. Zoetendal, E. G., Collier, C. T., Koike, S., Am. J. Gastroenterol. 85, 1307–1312 (1990). & Delzenne, N. M. Oligofructose promotes
Mackie, R. I. & Gaskins, H. R. Molecular 91. Alhabeeb, H., Chambers, E., Frost, G., satiety in rats fed a high-fat diet: involvement of
ecological analysis of the gastrointestinal Morrison, D. & Preston, T. Inulin propionate ester glucagon-like peptide-1. Obesity Res. 13,
microbiota: a review. J. Nutr. 134, 465–472 increases satiety and decreases appetite but 1000–1007 (2005).
(2004). does not affect gastric emptying in healthy 109. Zhou, J. et al. Peptide YY and proglucagon mRNA
74. Muyzer, G., De Waal, E. C. & Uitterlinden, A. G. humans. Proc. Nutr. Soc. 73, E21 (2014). expression patterns and regulation in the gut.
Profiling of complex microbial populations by 92. Chambers, E. S. et al. Effects of targeted delivery Obesity 14, 683–689 (2006).
denaturing gradient gel electrophoresis analysis of propionate to the human colon on appetite 110. Cani, P. D., Hoste, S., Guiot, Y. & Delzenne, N. M.
of polymerase chain reaction-amplified genes regulation, body weight maintenance and Dietary non-digestible carbohydrates promote
coding for 16S rRNA. Appl. Environ. Microbiol. 59, adiposity in overweight adults. Gut http:// L‑cell differentiation in the proximal colon of rats.
695–700 (1993). dx.doi.org/10.1136/gutjnl-2014-307913. Br. J. Nutr. 98, 32–37 (2007).
75. Zhao, L. The gut microbiota and obesity: from 93. Kondo, T., Kishi, M., Fushimi, T., Ugajin, S. 111. Pedersen, C. et al. Gut hormone release and
correlation to causality. Nat. Rev. Microbiol. 11, & Kaga, T. Vinegar intake reduces body weight, appetite regulation in healthy non-obese
639–647 (2013). body fat mass, and serum triglyceride levels in participants following oligofructose intake.
76. Hartstra, A. V., Bouter, K. E., Bäckhed, F. obese Japanese subjects. Biosci. Biotechnol. A dose-escalation study. Appetite 66, 44–53
& Nieuwdorp, M. Insights into the role of the Biochem. 73, 1837–1843 (2009). (2013).
microbiome in obesity and type 2 diabetes. 94. Theodorakis, M. J. et al. Human duodenal 112. Wichmann, A. et al. Microbial modulation of
Diabetes Care 38, 159–165 (2015). enteroendocrine cells: source of both incretin energy availability in the colon regulates
77. Turnbaugh, P. J. et al. A core gut microbiome in peptides, GLP‑1 and GIP. Am. J. Physiol. intestinal transit. Cell Host Microbe 14, 582–590
obese and lean twins. Nature 457, 480–484 Endocrinol. Metab. 290, E550–E559 (2006). (2013).
(2008). 95. Roberge, J. N. & Brubaker, P. L. Regulation of 113. Jiang, L. et al. Increased brain uptake and
78. Pedersen, M. H., Lauritzen, L. & Hellgren, L. I. intestinal proglucagon-derived peptide secretion oxidation of acetate in heavy drinkers. J. Clin.
Fish oil combined with SCFA synergistically by glucose-dependent insulinotropic peptide in a Invest. 123, 1605–1614 (2013).
prevent tissue accumulation of NEFA during novel enteroendocrine loop. Endocrinology 133, 114. Xiong, Y. et al. Short-chain fatty acids stimulate
weight loss in obese mice. Br. J. Nutr. 106, 233–240 (1993). leptin production in adipocytes through the
1449–1456 (2011). 96. De Silva, A. & Bloom, S. R. Gut hormones and G protein-coupled receptor GPR41. Proc. Natl
79. Ridaura, V. K. et al. Gut microbiota from twins appetite control: a focus on PYY and GLP‑1 as Acad. Sci. USA 101, 1045–1050 (2004).
discordant for obesity modulate metabolism in therapeutic targets in obesity. Gut Liver 6, 10–20 115. Soliman, M. et al. Inverse regulation of leptin
mice. Science 341, 1241214 (2013). (2012). mRNA expression by short-and long-chain fatty
80. Fujii, H. et al. Impact of dietary fiber intake on 97. Savage, A., Adrian, T., Carolan, G., Chatterjee, V. acids in cultured bovine adipocytes.
glycemic control, cardiovascular risk factors and & Bloom, S. Effects of peptide YY (PYY) on Domest. Anim. Endocrinol. 33, 400–409 (2007).
chronic kidney disease in Japanese patients mouth to caecum intestinal transit time and 116. Lee, S. & Hossner, K. Coordinate regulation of
with type 2 diabetes mellitus: the Fukuoka on the rate of gastric emptying in healthy ovine adipose tissue gene expression by
Diabetes Registry. Nutr. J. 12, 159 (2013). volunteers. Gut 28, 166–170 (1987). propionate. J. Anim. Sci. 80, 2840–2849 (2002).

NATURE REVIEWS | ENDOCRINOLOGY ADVANCE ONLINE PUBLICATION | 13


© 2015 Macmillan Publishers Limited. All rights reserved
REVIEWS

117. Ioffe, E., Moon, B., Connolly, E. & Friedman, J. M. in the white adipose tissue of high-fat diet-fed 155. Muccioli, G. G. et al. The endocannabinoid
Abnormal regulation of the leptin gene in the mice. J. Nutr. Biochem. 22, 712–722 (2011). system links gut microbiota to adipogenesis.
pathogenesis of obesity. Proc. Natl Acad. Sci. USA 136. Dewulf, E. M. et al. Evaluation of the relationship Mol. Syst. Biol. 6, 392 (2010).
95, 11852–11857 (1998). between GPR43 and adiposity in human. 156. Tedelind, S., Westberg, F., Kjerrulf, M. & Vidal, A.
118. Maffei, M. et al. Leptin levels in human and Nutr. Metab. 10, 11 (2013). Anti-inflammatory properties of the short-chain
rodent: measurement of plasma leptin and ob 137. Goossens, G. H. The role of adipose tissue fatty acids acetate and propionate: a study with
RNA in obese and weight-reduced subjects. dysfunction in the pathogenesis of obesity- relevance to inflammatory bowel disease.
Nat. Med. 1, 1155–1161 (1995). related insulin resistance. Physiol. Behav. 94, World J. Gastroenterol. 13, 2826–2832 (2007).
119. Frederich, R. C. et al. Leptin levels reflect body 206–218 (2008). 157. Liu, T. et al. Short-chain fatty acids suppress
lipid content in mice: evidence for diet-induced 138. Tilg, H. & Moschen, A. R. Adipocytokines: lipopolysaccharide-induced production of nitric
resistance to leptin action. Nat. Med. 1, mediators linking adipose tissue, inflammation oxide and proinflammatory cytokines through
1311–1314 (1995). and immunity. Nat. Rev. Immunol. 6, 772–783 inhibition of NF‑κB pathway in RAW264.7 cells.
120. Kondo, T., Kishi, M., Fushimi, T. & Kaga, T. (2006). Inflammation 35, 1676–1684 (2012).
Acetic acid upregulates the expression of genes 139. Apovian, C. M. et al. Adipose macrophage 158. Cox, M. A. et al. Short-chain fatty acids act as
for fatty acid oxidation enzymes in liver to infiltration is associated with insulin resistance antiinflammatory mediators by regulating
suppress body fat accumulation. J. Agric. Food and vascular endothelial dysfunction in obese prostaglandin E2 and cytokines. World J.
Chem. 57, 5982–5986 (2009). subjects. Arterioscler. Thromb. Vasc. Biol. 28, Gastroenterol. 15, 5549–5557 (2009).
121. Kimura, I. et al. Short-chain fatty acids and 1654–1659 (2008). 159. Ohira, H. et al. Butyrate attenuates inflammation
ketones directly regulate sympathetic nervous 140. Chaudhry, A. & Rudensky, A. Y. Control of and lipolysis generated by the interaction of
system via G protein-coupled receptor 41 inflammation by integration of environmental adipocytes and macrophages. J. Atheroscler.
(GPR41). Proc. Natl Acad. Sci. USA 108, cues by regulatory T cells. J. Clin. Invest. 123, Thromb. 20, 425–442 (2013).
8030–8035 (2011). 939–944 (2013). 160. Koves, T. R. et al. Mitochondrial overload and
122. Mithieux, G. & Gautier-Stein, A. Intestinal 141. Nishimura, S. et al. CD8+ effector T cells incomplete fatty acid oxidation contribute to
glucose metabolism revisited. Diabetes Res. contribute to macrophage recruitment and skeletal muscle insulin resistance. Cell Metab. 7,
Clin. Pract. 105, 295–301 (2014). adipose tissue inflammation in obesity. 45–56 (2008).
123. Ge, H. et al. Activation of G protein-coupled Nat. Med. 15, 914–920 (2009). 161. Corpeleijn, E., Saris, W. & Blaak, E. Metabolic
receptor 43 in adipocytes leads to inhibition of 142. Mazurek, T. et al. Human epicardial adipose flexibility in the development of insulin resistance
lipolysis and suppression of plasma free fatty tissue is a source of inflammatory mediators. and type 2 diabetes: effects of lifestyle.
acids. Endocrinology 149, 4519–4526 (2008). Circulation 108, 2460–2466 (2003). Obes. Rev. 10, 178–193 (2009).
124. Aberdein, N., Schweizer, M. & Ball, D. Sodium 143. Smith, P. M. et al. The microbial metabolites, 162. Yamashita, H. et al. Effects of acetate on lipid
acetate decreases phosphorylation of hormone short-chain fatty acids, regulate colonic Treg cell metabolism in muscles and adipose tissues of
sensitive lipase in isoproterenol-stimulated homeostasis. Science 341, 569–573 (2013). type 2 diabetic Otsuka Long-Evans Tokushima
3T3‑L1 mature adipocytes. Adipocyte 3, 144. Furusawa, Y. et al. Commensal microbe-derived Fatty (OLETF) rats. Biosci. Biotechnol. Biochem.
121–125 (2014). butyrate induces the differentiation of colonic 73, 570–576 (2009).
125. Rumberger, J. M, Arch, J. R. & Green, A. regulatory T cells. Nature 504, 446–450 (2013). 163. Fushimi, T. et al. Acetic acid feeding enhances
Butyrate and other short-chain fatty acids 145. Arpaia, N. et al. Metabolites produced by glycogen repletion in liver and skeletal muscle of
increase the rate of lipolysis in 3T3‑L1 commensal bacteria promote peripheral rats. J. Nutr. 131, 1973–1977 (2001).
adipocytes. PeerJ 2, e611 (2013). regulatory T‑cell generation. Nature 504, 164. Bonini, J. A., Anderson, S. M. & Steiner, D. F.
126. Hong, Y. H. et al. Acetate and propionate short 451–455 (2013). Molecular cloning and tissue expression of a
chain fatty acids stimulate adipogenesis via 146. Mariadason, J. M., Barkla, D. H. & Gibson, P. R. novel orphan G protein-coupled receptor from
GPCR43. Endocrinology 146, 5092–5099 Effect of short-chain fatty acids on paracellular rat lung. Biochem. Biophys. Res. Comm. 234,
(2005). permeability in Caco‑2 intestinal epithelium 190–193 (1997).
127. Kimura, I. et al. The gut microbiota suppresses model. Am. J. Physiol. 272, G705–G712 (1997). 165. Chai, W. et al. Glucagon-like peptide 1 recruits
insulin-mediated fat accumulation via the short- 147. Malago, J. et al. Differential modulation of microvasculature and increases glucose use in
chain fatty acid receptor GPR43. Nat. Commun. enterocyte-like Caco‑2 cells after exposure to muscle via a nitric oxide–dependent mechanism.
4, 1829 (2013). short-chain fatty acids. Food Addit. Contam. 20, Diabetes 61, 888–896 (2012).
128. Inoue, D., Tsujimoto, G. & Kimura, I. 427–437 (2003). 166. Cherbut, C. et al. Short-chain fatty acids modify
Regulation of energy homeostasis by GPR41. 148. Suzuki, T., Yoshida, S. & Hara, H. Physiological colonic motility through nerves and polypeptide
Front. Endocrinol. (Laussane) 5, 81 (2014). concentrations of short-chain fatty acids YY release in the rat. Am. J. Physiol. 275,
129. Zaibi, M. S. et al. Roles of GPR41 and GPR43 in immediately suppress colonic epithelial G1415–G1422 (1998).
leptin secretory responses of murine adipocytes permeability. Br. J. Nutr. 100, 297–305 (2008). 167. Tazoe, H. et al. Roles of short-chain fatty acids
to short chain fatty acids. FEBS Lett. 584, 149. Elamin, E. E., Masclee, A. A., Dekker, J., receptors, GPR41 and GPR43 on colonic
2381–2386 (2010). Pieters, H.‑J. & Jonkers, D. M. Short-chain fatty functions. J. Physiol. Pharmacol. 59, 251–262
130. Frost, G. et al. Effect of short chain fatty acids acids activate AMP-activated protein kinase and (2008).
on the expression of free fatty acid receptor 2 ameliorate ethanol-induced intestinal barrier 168. Samuel, B. S. et al. Effects of the gut microbiota
(Ffar2), Ffar3 and early-stage adipogenesis. dysfunction in Caco‑2 cell monolayers. J. Nutr. on host adiposity are modulated by the short-
Nutr. Diabetes 4, e128 (2014). 143, 1872–1881 (2013). chain fatty-acid binding G protein-coupled
131. Merkel, M., Eckel, R. H. & Goldberg, I. J. 150. Macia, L. et al. Metabolite-sensing receptors receptor, Gpr41. Proc. Natl Acad. Sci. USA 105,
Lipoprotein lipase genetics, lipid uptake, and GPR43 and GPR109A facilitate dietary fibre- 16767–16772 (2008).
regulation. J. Lipid Res. 43, 1997–2006 (2002). induced gut homeostasis through regulation 169. den Besten, G. et al. Gut-derived short-chain fatty
132. Grootaert, C. et al. Bacterial monocultures, of the inflammasome. Nat. Commun. 6, 6734 acids are vividly assimilated into host
propionate, butyrate and H2O2 modulate the (2015). carbohydrates and lipids. Am. J. Physiol. 305,
expression, secretion and structure of the 151. Cani, P. D. et al. Selective increases of G900–G910 (2013).
fasting induced adipose factor in gut epithelial bifidobacteria in gut microflora improve 170. Demigne, C. et al. Effect of propionate on fatty
cell lines. Environ. Microbiol. 13, 1778–1789 high‑fat‑diet-induced diabetes in mice through acid and cholesterol synthesis and on acetate
(2011). a mechanism associated with endotoxaemia. metabolism in isolated rat hepatocytes. Br. J.
133. Haberland, M., Carrer, M., Mokalled, M. H., Diabetologia 50, 2374–2383 (2007). Nutr. 74, 209–219 (1995).
Montgomery, R. L. & Olson, E. N. Redundant 152. Cani, P. D. et al. Metabolic endotoxemia initiates 171. Hardie, D. & Pan, D. Regulation of fatty acid
control of adipogenesis by histone deacetylases obesity and insulin resistance. Diabetes 56, synthesis and oxidation by the AMP-activated
1 and 2. J. Biol. Chem. 285, 14663–14670 1761–1772 (2007). protein kinase. Biochem. Soc. Trans. 30,
(2010). 153. Mehta, N. N. et al. Experimental endotoxemia 1064–1070 (2002).
134. Li, G., Yao, W. & Jiang, H. Short-chain fatty acids induces adipose inflammation and insulin 172. Zhang, B. B., Zhou, G. & Li, C. AMPK: an emerging
enhance adipocyte differentiation in the stromal resistance in humans. Diabetes 59, 172–181 drug target for diabetes and the metabolic
vascular fraction of porcine adipose tissue. (2010). syndrome. Cell Metab. 9, 407–416 (2009).
J. Nutr. 144, 1887–1895 (2014). 154. Vila, I. K. et al. Immune cell toll-like receptor 4 173. Li, X. et al. Acetic acid activates the AMP-activated
135. Dewulf, E. M. et al. Inulin-type fructans with mediates the development of obesity-and protein kinase signaling pathway to regulate lipid
prebiotic properties counteract GPR43 endotoxemia-associated adipose tissue fibrosis. metabolism in bovine hepatocytes. PLoS ONE 8,
overexpression and PPARγ-related adipogenesis Cell Rep. 7, 1116–1129 (2014). e67880 (2013).

14 | ADVANCE ONLINE PUBLICATION www.nature.com/nrendo


© 2015 Macmillan Publishers Limited. All rights reserved
REVIEWS

174. Yamashita, H. et al. Improvement of obesity and probiotics for the gut-liver axis. PLoS ONE 8, 180. Venter, C. S., Vorster, H. H. & Cummings, J. H.
glucose tolerance by acetate in type 2 diabetic e63388 (2013). Effects of dietary propionate on
Otsuka Long-Evans Tokushima Fatty (OLETF) 177. Wolever, T., Spadafora, P. & Eshuis, H. carbohydrate and lipid metabolism in healthy
rats. Biosci. Biotechnol. Biochem. 71, Interaction between colonic acetate and volunteers. Am. J. Gastroenterol. 85, 549–553
1236–1243 (2007). propionate in humans. Am. J. Clin. Nutr. 53, (1990).
175. Sakakibara, S., Yamauchi, T., Oshima, Y., 681–687 (1991).
Tsukamoto, Y. & Kadowaki, T. Acetic acid 178. Laurent, C. et al. Effect of acetate and Acknowledgements
activates hepatic AMPK and reduces propionate on fasting hepatic glucose Research in the authors’ laboratory is funded by
hyperglycemia in diabetic KK‑A(y) mice. production in humans. Eur. J. Clin. Nutr. 49, TI Food and Nutrition, a public–private partnership
Biochem. Biophys. Res. Comm. 344, 597–604 484–491 (1995). on precompetitive research in food and nutrition.
(2006). 179. Fernandes, J., Vogt, J. & Wolever, T. M.
176. Endo, H., Niioka, M., Kobayashi, N., Tanaka, M. Intravenous acetate elicits a greater free fatty Author contributions
& Watanabe, T. Butyrate-producing probiotics acid rebound in normal than hyperinsulinaemic All authors researched the data for the article,
reduce nonalcoholic fatty liver disease humans. Eur. J. Clin. Nutr. 66, 1029–1034 discussed the content, wrote, edited and approved
progression in rats: new insight into the (2012). the manuscript before submission.

NATURE REVIEWS | ENDOCRINOLOGY ADVANCE ONLINE PUBLICATION | 15


© 2015 Macmillan Publishers Limited. All rights reserved

Você também pode gostar