ArticlePDF Available

Prolonged Fasting Identifies Skeletal Muscle Mitochondrial Dysfunction as Consequence Rather Than Cause of Human Insulin Resistance

Authors:

Abstract and Figures

Type 2 diabetes and insulin resistance have been associated with mitochondrial dysfunction, but it is debated whether this is a primary factor in the pathogenesis of the disease. To test the concept that mitochondrial dysfunction is secondary to the development of insulin resistance, we employed the unique model of prolonged fasting in humans. Prolonged fasting is a physiologic condition in which muscular insulin resistance develops in the presence of increased free fatty acid (FFA) levels, increased fat oxidation and low glucose and insulin levels. It is therefore anticipated that skeletal muscle mitochondrial function is maintained to accommodate increased fat oxidation unless factors secondary to insulin resistance exert negative effects on mitochondrial function. While in a respiration chamber, twelve healthy males were subjected to a 60 h fast and a 60 h normal fed condition in a randomized crossover design. Afterward, insulin sensitivity was assessed using a hyperinsulinemic-euglycemic clamp, and mitochondrial function was quantified ex vivo in permeabilized muscle fibers using high-resolution respirometry. Indeed, FFA levels were increased approximately ninefold after 60 h of fasting in healthy male subjects, leading to elevated intramuscular lipid levels and decreased muscular insulin sensitivity. Despite an increase in whole-body fat oxidation, we observed an overall reduction in both coupled state 3 respiration and maximally uncoupled respiration in permeabilized skeletal muscle fibers, which could not be explained by changes in mitochondrial density. These findings confirm that the insulin-resistant state has secondary negative effects on mitochondrial function. Given the low insulin and glucose levels after prolonged fasting, hyperglycemia and insulin action per se can be excluded as underlying mechanisms, pointing toward elevated plasma FFA and/or intramuscular fat accumulation as possible causes for the observed reduction in mitochondrial capacity.
Content may be subject to copyright.
Prolonged Fasting Identifies Skeletal Muscle
Mitochondrial Dysfunction as Consequence Rather Than
Cause of Human Insulin Resistance
Joris Hoeks,
1
Noud A. van Herpen,
1,2
Marco Mensink,
3
Esther Moonen-Kornips,
1
Denis van Beurden,
1,2
Matthijs K.C. Hesselink,
4
and Patrick Schrauwen
1,2
OBJECTIVE—Type 2 diabetes and insulin resistance have been
associated with mitochondrial dysfunction, but it is debated
whether this is a primary factor in the pathogenesis of the
disease. To test the concept that mitochondrial dysfunction is
secondary to the development of insulin resistance, we employed
the unique model of prolonged fasting in humans. Prolonged
fasting is a physiologic condition in which muscular insulin
resistance develops in the presence of increased free fatty acid
(FFA) levels, increased fat oxidation and low glucose and insulin
levels. It is therefore anticipated that skeletal muscle mitochon-
drial function is maintained to accommodate increased fat oxi-
dation unless factors secondary to insulin resistance exert
negative effects on mitochondrial function.
RESEARCH DESIGN AND METHODS—While in a respiration
chamber, twelve healthy males were subjected to a 60 h fast and
a 60 h normal fed condition in a randomized crossover design.
Afterward, insulin sensitivity was assessed using a hyperinsuline-
mic-euglycemic clamp, and mitochondrial function was quanti-
fied ex vivo in permeabilized muscle fibers using high-resolution
respirometry.
RESULTS—Indeed, FFA levels were increased approximately
ninefold after 60 h of fasting in healthy male subjects, leading to
elevated intramuscular lipid levels and decreased muscular insu-
lin sensitivity. Despite an increase in whole-body fat oxidation,
we observed an overall reduction in both coupled state 3
respiration and maximally uncoupled respiration in permeabil-
ized skeletal muscle fibers, which could not be explained by
changes in mitochondrial density.
CONCLUSIONS—These findings confirm that the insulin-resis-
tant state has secondary negative effects on mitochondrial func-
tion. Given the low insulin and glucose levels after prolonged
fasting, hyperglycemia and insulin action per se can be excluded
as underlying mechanisms, pointing toward elevated plasma FFA
and/or intramuscular fat accumulation as possible causes for the
observed reduction in mitochondrial capacity. Diabetes 59:
2117–2125, 2010
Although the existence of mitochondrial abnor-
malities in patients with type 2 diabetes has
been extensively reported during the last de-
cade (1–5), there is no evidence that a reduced
mitochondrial function is a primary factor in the patho-
physiology of this disease. In fact, alternative theories
state that impaired mitochondrial capacity is secondary to
the insulin-resistant or diabetic state. In this context, it has
been shown that insulin can stimulate mitochondrial bio-
genesis and increases ATP synthesis in skeletal muscle
(6,7). A reduced insulin action in skeletal muscle, as
observed in type 2 diabetic patients, could therefore
contribute to the origin of mitochondrial dysfunction.
Additionally, the increased exposure of skeletal muscle
mitochondria to elevated levels of free fatty acids (FFA),
seen in insulin resistance and type 2 diabetes, has been
suggested to interfere with proper mitochondrial function.
Thus, Szendroedi et al. (8) showed that plasma FFA levels
negatively correlated with mitochondrial function mea-
sured by magnetic resonance spectroscopy. Furthermore,
we showed that the acute elevation of plasma FFA by lipid
infusion is accompanied by downregulation of the tran-
scriptional coactivator peroxisome proliferator-activated
receptor gamma coactivator-1(PGC1) and other genes
involved in mitochondrial metabolism (9). Moreover, it
was shown in a comparable study that short-term eleva-
tion of lipid availability reduces insulin-stimulated in-
crease in ATP synthase flux in skeletal muscle (10),
although this may mainly reflect an effect of muscular
insulin resistance on ATP flux.
Prolonged fasting (48 h) in humans is accompanied by
a reduction in insulin sensitivity, elevated plasma FFA
levels, elevated intramuscular fat levels, but also an in-
crease in whole-body fat oxidative capacity (11,12). Fur-
thermore, prolonged fasting-induced insulin resistance is
not accompanied by hyperglycemia or hyperinsulinemia,
factors that have been suggested to cause mitochondrial
dysfunction in diabetes (7,13). In fact, prolonged fasting is
a physiologic condition in which insulin resistance devel-
ops to spare glucose for utilization by the brain, and
increased FFA levels are accompanied by increased fat
oxidation. It could therefore be anticipated that despite
the development of insulin resistance, mitochondrial func-
tion is maintained to accommodate increased fat oxidation
during prolonged fasting. Alternatively, if (lipid-induced)
insulin resistance or factors associated with the insulin-
resistant state indeed cause mitochondrial dysfunction,
we anticipate a reduction in mitochondrial function with
prolonged fasting. Therefore, we aim to test the concept
that mitochondrial dysfunction originates secondary to
From the
1
School for Nutrition, Toxicology and Metabolism, Department of
Human Biology, Maastricht University Medical Centre, Maastricht, the
Netherlands;
2
Top Institute Food and Nutrition, Wageningen, the Nether-
lands; the
3
Department of Human Nutrition, Wageningen University,
Wageningen, the Netherlands; and the
4
School for Nutrition, Toxicology and
Metabolism, Department of Human Movement Sciences, Maastricht Univer-
sity Medical Centre, Maastricht, the Netherlands.
Corresponding author: J. Hoeks, j.hoeks@hb.unimaas.nl.
Received 13 April 2010 and accepted 15 June 2010. Published ahead of print
at http://diabetes.diabetesjournals.org on 23 June 2010. DOI: 10.2337/db10-0519.
J.H. and N.A.v.H. contributed equally to this study.
The study has been registered at www.trialregister.nl with registration num-
ber NTR 2042.
© 2010 by the American Diabetes Association. Readers may use this article as
long as the work is properly cited, the use is educational and not for profit,
and the work is not altered. See http://creativecommons.org/licenses/by
-nc-nd/3.0/ for details.
The costs of publication of this article were defrayed in part by the payment of page
charges. This article must therefore be hereby marked “advertisement” in accordance
with 18 U.S.C. Section 1734 solely to indicate this fact.
ORIGINAL ARTICLE
diabetes.diabetesjournals.org DIABETES, VOL. 59, SEPTEMBER 2010 2117
the development of insulin resistance by employing the
physiologic model of prolonged fasting-induced insulin
resistance.
RESEARCH DESIGN AND METHODS
Twelve healthy, lean, male volunteers who had no family history of diabetes
or any other endocrine disorder participated in this study (Table 1). None of
the subjects engaged in sports activities for more than 2 h per week. Body
composition (14) and maximal aerobic capacity (15) were measured as
described previously. The study protocol was reviewed and approved by the
Medical Ethical Committee of Maastricht University Medical Centre and all
subjects gave their written informed consent before participating in the study.
Experimental design. Subjects participated in two experimental trials: a
60 h fast and a 60 h normal fed condition, in a randomized crossover design
with a 2-week washout period. In the fast condition, subjects were fasted for
60 h (calorie-free drinks only), whereas in the second condition, subjects were
fed in energy balance (50 –35–15% of energy as carbohydrates, fat, and protein,
respectively). Before the start of each experimental period, a standardized
evening meal was provided. Subject stayed in a respiration chamber during
the entire 60 h to ensure compliance to the dietary regime and to allow the
measurement of 24 h substrate oxidation and energy expenditure (16). In the
respiration chamber, subjects followed an activity protocol as previously
described (17). During the intervention, blood samples were taken after 12, 36,
and 60 h after an overnight fast in case of the fed condition.
Hyperinsulinemic-euglycemic clamp. After leaving the respiration chamber
on the morning of the third day, a muscle biopsy was taken (18) and a
hyperinsulinemic-euglycemic clamp procedure (4) was performed. Insulin-
stimulated plasma glucose rate of disappearance (Rd), endogenous glucose
production (EGP), and nonoxidative glucose disposal (mainly reflecting
glycogen synthesis) were calculated as by Phielix et al. (4). Substrate
oxidation in the basal and the insulin-stimulated state was measured using
indirect calorimetry (Omnical, Maastricht, the Netherlands) and calculated
according to Frayn (19).
Muscle biopsy. After taking the muscle biopsy, a portion of the muscle tissue
was directly frozen in melting isopentane and stored at 80°C until assayed.
Another portion (30 mg) was immediately placed in ice cold preservation
medium (4).
Blood analyses. Plasma nonesterified fatty acids (Wako Nefa C test kit; Wako
Chemicals, Neuss, Germany) and glucose (hexokinase method; LaRoche,
Basel, Switzerland) were measured with enzymatic assays automated on a
Cobas Fara/Mira. Insulin concentration was determined using a radioimmu-
noassay (Linco Reseach, St. Charles, MO).
Intramuscular triacylglycerols. Fresh cryosections (5 m) were stained for
intramuscular triacylglycerols (IMTG) by Oil Red O staining combined with
fibertyping and immunolabeling of the basal membrane marker laminin to
allow quantification of IMTG, as described previously (20,21).
Mitochondrial DNA copy number and citrate synthase activity. Mito-
chondrial DNA (mtDNA) copy number, the ratio of NADH dehydrogenase
subunit one (ND1) to lipoprotein lipase (LPL) (mtDNA/nuclear DNA) was
determined as described previously (4). Citrate synthase (CS) activity was
measured spectrophotometrically as described previously (22).
High resolution respirometry. Permeabilized skeletal muscle fibers were
immediately prepared from the muscle tissue collected in the preservation
medium, as described elsewhere (4,23). Subsequently, the permeabilized
muscle fibers (2.5 mg wet weight) were analyzed for mitochondrial function
using an oxygraph (OROBOROS Instruments, Innsbruck, Austria), in essence
according to Phielix et al. (4). To prevent oxygen limitation, the respiration
chambers were hyperoxygenated up to 500 mol/l O
2
. Subsequently, two
different multisubstrate/inhibition protocols were used in which substrates
and inhibitors were added consecutively in saturating concentrations. State 2
respiration was measured after the addition of malate (4 mmol/l) plus
octanoyl-carnitine (50 mol/l) or malate (4 mmol/l) plus glutamate (10
mmol/l). Subsequently, an excess of 2 mmol/l of ADP was added to determine
coupled (state 3) respiration. Coupled respiration was then maximized with
convergent electron input through Complex I and Complex II by adding
saturating concentrations of succinate (10 mmol/l). Finally, the chemical
uncoupler carbonylcyanide-4-(trifluoromethoxy)-phenylhydrazone (FCCP)
was titrated or oligomycin (2 g/ml) was added to evaluate the maximal
capacity of the electron transport chain and the respiration not coupled to
ATP synthesis (state 4o respiration), respectively. The integrity of the outer
mitochondrial membrane was assessed by the addition of cytochrome C (10
mol/l) upon maximal coupled respiration. All measurements were performed
in duplicate.
Western blotting. Oxidative phosphorylation (OXPHOS) protein levels and
mitochondrial uncoupling protein-3 (UCP3) content were measured in whole
muscle by Western blotting as described previously (24). UCP3 was expressed
as a ratio over the sum of OXPHOS complexes to correct for differences in
mitochondrial density.
Statistics. Data are reported as means SE. Statistical analyses were
performed using the statistical computer program SPSS 16.0 for Mac OS X.
Statistical comparisons between the two conditions (fed versus fast) were
performed using the paired Student ttest. Plasma FFA, glucose, and insulin
were compared over time by two-way repeated measures ANOVA for inves-
tigation of treatment and time (treatment*time) interactions. When the
interaction was significant, we performed post hoc testing to determine the
exact location of the difference. Differences were considered statistically
significant if P0.05.
RESULTS
Plasma parameters. A significant treatment*time inter-
action (P0.001) was observed for all plasma parameters
(Fig. 1). At the start of the intervention (t 12 h, after an
overnight fast) plasma FFA levels (Fig. 1A) were similar
between both conditions (221 18 vs. 215 27 mol/l,
respectively, P0.82). Upon 60 h of fasting, plasma FFA
increased dramatically, up to 1,981 95 mol/l vs. 387
37 in the fed condition (P0.001).
Plasma glucose values (Fig. 1B) were similar at baseline,
averaging 5.05 0.09 and 4.98 0.06 mmol/l in the fed and
the fasted state, respectively (P0.28), and remained
unchanged throughout the fed condition. During fasting,
however, plasma glucose levels gradually decreased to
3.72 0.13 mmol/l at t 60h(P0.001).
Baseline plasma insulin levels (Fig. 1C) were similar in
both conditions (12.9 0.9 vs. 11.9 1.2 U/ml in fed
versus fasted, respectively, P0.25) and did not change in
the fed condition. In the fasted condition however, plasma
insulin levels were markedly reduced to 7.1 0.6 U/ml at
t36h(P0.001) and were maintained at this lower
level (7.0 U/ml 0.74) at t 60h(P0.001).
Indirect calorimetry. Twenty-four hour energy expendi-
ture during the last 24 h of the 60 h intervention was
slightly but significantly reduced upon prolonged fasting
(10.88 0.33 vs. 10.30 0.30 MJ/day, in fed versus fasted,
respectively, P0.02). The difference was mainly caused
by a reduction in diet-induced thermogenesis, and not
caused by a decrease in resting metabolic rate (data not
shown). Additionally, whole-body 24-h fat oxidation was
increased upon prolonged fasting as evidenced by a sig-
nificant reduction in 24-h respiratory exchange ratio
(RER) (0.91 0.009 vs. 0.77 0.003 in fed versus fasted,
respectively, P0.001).
Insulin sensitivity. All subjects displayed a decrease in
glucose infusion rate upon 60 h of fasting (Fig. 2A). We
also calculated the insulin sensitivity index, an index that
takes into account the variation in insulin and glucose
levels during the clamp (25). Insulin sensitivity index was
reduced by 45% upon 60 h of fasting, as compared with
the fed condition (Fig. 2B,P0.001).
The reduction in whole-body insulin sensitivity was
mainly accounted for by a reduction in insulin-stimulated
TABLE 1
Subject characteristics
Parameter Mean SE
Age (years) 23.6 1.0
Body weight (kg) 78.5 2.5
Fat-free mass (kg) 65.9 1.8
Height (m) 1.86 0.02
BMI (kg/m
2
)22.6 0.5
Maximal aerobic capacity (ml O
2
/kg
FFM
/min) 57.5 1.5
PROLONGED FASTING AND MITOCHONDRIAL FUNCTION
2118 DIABETES, VOL. 59, SEPTEMBER 2010 diabetes.diabetesjournals.org
glucose disposal (Rd, P0.001, Table 2). The reduced
insulin-stimulated glucose disposal after fasting, mainly
reflecting muscle glucose uptake, was due to both reduced
insulin-stimulation of glucose oxidation and reduced non-
oxidative glucose disposal (Table 2). However, insulin-
stimulated glucose oxidation seemed to be more severely
suppressed by 60 h of fasting (Table 2). Also, baseline
endogenous glucose production was reduced after 60 h of
fasting. However, insulin-induced suppression of endoge-
nous glucose production, reflecting hepatic insulin sensi-
tivity, was only marginally affected by fasting and was
almost complete in both conditions (Table 2).
Metabolic flexibility. Metabolic flexibility was blunted in
the fasted condition when compared with the fed condi-
tion (Fig. 3A,P0.001). Basal whole-body fat oxidation
was increased by 1.5-fold upon prolonged fasting (Fig. 3B,
P0.001). During the glucose clamp, fat oxidation
significantly decreased in both conditions (P0.001), but
the suppression was significantly less in the fasted condi-
tion (Fig. 3B,P0.001). Basal carbohydrate oxidation
after fasting was only 35% of the value obtained in the
fed situation (P0.001), but increased in both conditions
during the glucose clamp (Fig. 3C). However, this insulin-
induced change in carbohydrate oxidation was blunted
upon fasting (P0.001).
Intramuscular triacylglycerols. The mean intramuscu-
lar IMTG area fraction (Fig. 4), was 2.7-fold higher after
60 h of fasting in comparison with the fed condition (P
0.001). The increase in lipid accumulation was more
pronounced (3.5 fold, P0.001) in fibers identified as
slow, oxidative (type 1) fibers. Within type 2 muscle fibers,
IMTG levels increased by approximately twofold after 60 h
of fasting (P0.015).
Mitochondrial function. Mitochondrial DNA copy num-
ber was similar in both the fed and the fasting condition (Fig.
5A,P0.96). Also, OXPHOS protein levels (Complex I:
21.2 6.4 vs. 17.7 5.7 arbitrary units, P0.50; Complex II:
48.7 14.6 vs. 48.3 16.2 AU, P0.97; Complex III: 11.6
1.5 vs. 11.9 1.1 AU, P0.88; Complex IV: 92.8 7.0 vs.
93.7 9.1 AU, P0.85; Complex V: 4.4 0.7 vs. 4.8 0.9,
P0.66) and CS activity (76.2 7.3 vs. 70.2 6.6
mol/min/g protein, P0.41) were similar in the fed versus
the fasted state, respectively, confirming an equal mitochon-
drial mass in both conditions.
Nonetheless, we adjusted the oxygen fluxes for individ-
ual differences in mtDNA copy number. However, similar
results were obtained without this correction (see supple-
mental Fig. 1 in the online appendix available at http://
diabetes.diabetesjournals.org). State 2 respiration (i.e.,
respiration in the presence of substrate alone) was not
different between conditions on any of the substrate
combinations studied (Fig. 5B). However, ADP-stimulated
(state 3) respiration on a lipid substrate (malate oc-
tanoyl-carnitine, MO) was significantly reduced upon fast-
ing, as compared with the fed situation (P0.03, Fig. 5C).
Similarly, state 3 upon the Complex I substrates malate
glutamate (MG) was 22% lower after fasting, although
this did not reach statistical significance (P0.12, Fig.
5D).
Additionally, respiration upon parallel electron input to
both Complex I and II was reduced by 20% upon
prolonged fasting. Thus, state 3 respiration upon malate
octanoyl-carnitine glutamate (MOG) was significantly
lower in the fasted state compared with the fed condition
(P0.01, Fig. 5E). Similar differences were observed for
state 3 respiration upon malate octanoyl-carnitine
glutamate succinate (MOGS, P0.02, Fig. 5E) and
malate glutamate succinate (MGS, P0.04, Fig. 5E).
Maximal FCCP-induced uncoupled respiration, reflecting
the maximal capacity of the electron transport chain, was
also reduced by 23% after 60 h of fasting (P0.008, Fig.
5F). Finally, state 4o respiration (reflecting mitochondrial
proton leak) was similar between the fed and fasted state
(P0.48, Fig. 5G). The average increase in oxygen
consumption upon cytochrome C was less than 10% (un-
Time (hours)
12 36 60
Plasma Glucose (mmol/L)
0
1
2
3
4
5
6
Time (hours)
12 36 60
Plasma FFA (mmol/L)
0.0
0.5
1.0
1.5
2.0
2.5
*
*
**
Time (hours)
12 36 60
Plasma Insulin (µU/ml)
0
2
4
6
8
10
12
14
16
**
A
B
C
FIG. 1. Plasma free fatty acids (A), plasma glucose (B), and plasma
insulin (C) levels after 12, 36, and 60 h of fasting. Open circles
represent the fed condition; closed circles represent the fasted condi-
tion. Values are mean SE. *P<0.05.
J. HOEKS AND ASSOCIATES
diabetes.diabetesjournals.org DIABETES, VOL. 59, SEPTEMBER 2010 2119
derscoring the viability of the muscle fibers) and similar in
both conditions (P0.54).
Mitochondrial UCP3 did not change upon 60 h of fasting
in humans and averaged 2.21 0.38 vs. 2.20 0.41 AU in
the fed versus fasted condition (P0.96).
DISCUSSION
In this study, we evaluated the effect of prolonged fasting
on skeletal muscle mitochondrial functional capacity in
humans to examine whether the mitochondrial dysfunc-
tion that is frequently reported in insulin resistance and
type 2 diabetes can be a consequence of lipid-induced
insulin resistance, rather than a cause. In contrast to the
hyperglycemia and hyperinsulinaemia accompanying
“energy excess”-induced insulin resistance (lipid infu-
sion, high-fat diets), prolonged fasting-induced insulin
resistance is associated with hypoglycemia and hypoin-
sulinemia. Moreover, prolonged fasting-induced lipid
accumulation and insulin resistance are considered to be a
functional physiologic response. Thus, reduced insulin
sensitivity saves carbohydrates for the central nervous
system, being obligate for glucose and not requiring insu-
lin for its uptake, whereas increased lipid availability at
the same time can serve as a direct available energy source
for the muscles and is paralleled by an enhanced fat
oxidative capacity (12). Therefore, we anticipated that
skeletal muscle mitochondrial function would not be
impaired in this model unless mitochondrial function is
impaired by factors that are secondary to the lipid-induced
insulin-resistant state. Intriguingly, we found that only 60 h
of fasting in humans was accompanied by an overall
reduction in skeletal muscle mitochondrial capacity,
which was not explained by changes in mitochondrial
density.
We assessed mitochondrial functional capacity in detail
(Fig. 5) by using a wide variety of substrates and substrate
combinations to determine the maximum ADP-stimulated
respiration (state 3) fueled by a lipid substrate, by Com-
plex I substrates, and upon parallel electron input into
Complex I (NADH) and II (FADH
2
). Interestingly, state 3
respiration was reduced by 20% upon all substrates,
which reduces the possibility that the decline is caused by
substrate-specific alterations such as substrate uptake into
the mitochondria. Therefore, both a reduction in the
activity of the electron transport chain and the oxidative
phosphorylation system could underlie the reduced state 3
respiration. Using the chemical uncoupler FCCP, control
over respiration by the oxidative phosphorylation system
is bypassed; thus, FCCP-induced respiration reflects the
maximal capacity of the electron transport chain. Irrespec-
tive of the intervention, FCCP was able to enhance mito-
chondrial respiration considerably over state 3 values,
indicating that the electron transport chain is not rate-
limiting in state 3. However, FCCP-induced respiration in
Fed Fasted
Glucose Infusion Rate (ml/h)
40
60
80
100
120
140
160
180
200
220
240
SI index (ml min-1 kg-1/µU ml-1)
0
2
4
6
8
10
12
*
AB
FIG. 2. Assessment of insulin sensitivity by hyperinsulinemic-euglycemic clamping after 60 h of fasting. Arepresents the individual data for the
glucose infusion rate. Bdisplays the group results of the insulin sensitivity index, i.e., the glucose infusion rate corrected for body weight,
glucose, and plasma insulin levels during the clamp procedure. The white bar represents the fed condition whereas the black bar depicts the fasted
condition. Values are mean SE. *P<0.05. S
I
, insulin sensitivity index.
TABLE 2
Substrate kinetics
Fed Fasted
Rd Glucose (mol/kg/min)
Basal 11.3 0.6 7.6 0.4*
Clamp 38.0 2.8 18.8 0.9*
Delta 26.7 2.6 11.2 1.0*
EGP (mol/kg/min)
Basal 10.7 0.6 6.8 0.3*
Clamp 1.0 0.4 0.7 0.4*
Delta absolute 11.7 0.6 6.07 0.5*
Delta % 111.6 5.9 90.0 6.7*
CHO oxidation (mol/kg/min)
Basal 9.7 0.8 3.4 0.6*
Clamp 20.8 1.1 6.6 0.8*
Delta 11.1 0.9 3.4 0.7*
NOGD (mol/kg/min)
Basal 1.6 0.5 4.2 0.8*
Clamp 17.2 2.2 12.6 0.8*
Delta 15.6 2.4 8.2 1.1*
Lipid oxidation (mol/kg/min)
Basal 1.2 0.0 1.8 0.2*
Clamp 0.3 0.1 1.52 0.1*
Delta 0.8 0.1 0.31 0.1*
Values are mean SE. *P0.05. EGP, endogenous glucose
production; CHO, carbohydrate; NOGD, nonoxidative glucose
disposal.
PROLONGED FASTING AND MITOCHONDRIAL FUNCTION
2120 DIABETES, VOL. 59, SEPTEMBER 2010 diabetes.diabetesjournals.org
itself was reduced by 23% upon prolonged fasting (Fig.
5F), indicating that a combined reduction of both the
capacity of the electron transport chain and oxidative
phosphorylation underlies the reduced mitochondrial ox-
idative capacity upon fasting.
The reduced mitochondrial capacity was not accounted
for by a reduction in mitochondrial density. Thus, mtDNA
copy number, CS activity, and OXPHOS protein levels
remained unaffected by fasting. Although this does not
exclude the possibility that prolonged exposure to high
FFA and/or insulin resistance may lead to a reduced
mitochondrial function as observed in type 2 diabetes via
reduced mitochondrial biogenesis (e.g., via PGC-1), this
finding indicates that fasting interferes with intrinsic mi-
tochondrial capacity in skeletal muscle. Interestingly, a
similar reduction in intrinsic mitochondrial capacity, with-
out differences in mitochondrial content, was recently
reported by us in type 2 diabetic patients and first degree
relatives when compared with BMI- and age-matched
obese control subjects (4). Thus, the reduction in mito-
chondrial function upon fasting mimics the situation as
observed in the diabetic state, and suggests that similar
(secondary) effects are involved in causing mitochondrial
dysfunction. However, it should be noted that fasting is
not a direct model for type 2 diabetes. Furthermore,
although the available data in the literature underpin the
notion that fasting-induced lipid accumulation is respon-
sible for reduced insulin sensitivity upon fasting, we
cannot exclude the possibility that the reduced mitochon-
drial function upon prolonged fasting triggers the insulin
resistance observed.
Resistance of skeletal muscle to insulin action per se
has been suggested to explain the reduction in mitochon-
drial functional capacity observed in diabetes. Thus, in
healthy individuals, it was shown thata7hinsulin infusion
increased mitochondrial protein synthesis, cytochrome C
oxidase (COX), and citrate synthase (CS) enzyme activi-
ties and ATP production (26). Moreover, it has been
reported that exposing human primary muscle cells to
insulin upregulates the expression of PGC-1(27). There-
fore, the reduction in insulin action in type 2 diabetes may
underlie the observed mitochondrial defects. In agreement
with this hypothesis, it was demonstrated that at low
doses of insulin (reflecting postabsorptive levels) the
skeletal mitochondrial ATP synthesis rate was not differ-
ent between diabetic patients and controls, indicating that
Basal Insulin-stimulated
Fat oxidation (µmol kg-1 min-1)
0.0
0.5
1.0
1.5
2.0
2.5
Basal Insulin-stimulated
Respiratory Exchange Ratio
0.70
0.75
0.80
0.85
0.90
0.95
1.00
*
*
A
Basal Insulin-stimulated
Carbohydrate oxidation (µmol kg-1 min-1)
0
5
10
15
20
25
*
*
B
C
*
*
FIG. 3. Indirect calorimetry results after 60 h of fasting in the basal
state and upon the hyperinsulinemic-euglycemic clamp. Ashows that
metabolic flexibility, defined as the change in respiratory exchange
ratio upon insulin stimulation, is blunted upon prolonged fasting. B
and Cdisplay whole-body lipid and carbohydrate oxidation, respec-
tively. White bars/circles represent the fed condition; black bars/circles
represent the fasted condition. Values are mean SE. *P<0.05.
Area fraction (%)
0
2
4
6
8
10
12
*
Tot a l Type 2Type 1
*
*
FIG. 4. Intramuscular triacylglycerols measured by Oil Red O staining
combined with an immunofluorescence staining against slow myosin
heavy chain (sMHC), to determine fibertyping. White bars represent
the fed condition; black bars represent the fasted condition. Values are
mean SE. *P<0.05.
J. HOEKS AND ASSOCIATES
diabetes.diabetesjournals.org DIABETES, VOL. 59, SEPTEMBER 2010 2121
State 3, fat oxidation
0
10
20
30
40
50
State 3, complex I+II
MOG MOGS MGS
0
20
40
60
80
100
120
140
Mitochondrial density
Mitochondrial DNA copy number (AU)
0
2000
4000
6000
8000
10000
A
State 2
MMOMG
O
2
flux ([pmol mg
-1
s
-1
] / mtDNA copy # *10
4
)
O
2
flux ([pmol mg
-1
s
-1
] / mtDNA copy # *10
4
)
O
2
flux ([pmol mg
-1
s
-1
] / mtDNA copy # *10
4
)
O
2
flux ([pmol mg
-1
s
-1
] / mtDNA copy # *10
4
)
O
2
flux ([pmol mg
-1
s
-1
] / mtDNA copy # *10
4
)
O
2
flux ([pmol mg
-1
s
-1
] / mtDNA copy # *10
4
)
0
2
4
6
8
10
12
14
B
CE
**
State uncoupled, complex I+II
0
20
40
60
80
100
120
140
160
180
200
220
F
*
State 4o, complex I+II
0
10
20
30
40
50
60
G
*
*
State 3, complex I
0
20
40
60
80
D
MO MG
MOGS MGS
FIG. 5. Evaluation of mitochondrial function. A: Mitochondrial density. B: Mitochondrial respiration upon substrates only (state 2). C:
ADP-stimulated respiration (state 3) upon a lipid substrate. D: State 3 respiration fueled by Complex I-linked substrates. E: State 3 respiration
upon parallel electron input into Complex I and II. F: Maximally uncoupled respiration upon FCCP. G: Mitochondrial respiration uncoupled from
ATP synthesis (state 4o). White bars represent the fed condition, black bars represent the fasted condition. Values are mean SE, n10. *P<
0.05. M, malate; O, octanoyl-carnitine; G, glutamate; S, succinate.
PROLONGED FASTING AND MITOCHONDRIAL FUNCTION
2122 DIABETES, VOL. 59, SEPTEMBER 2010 diabetes.diabetesjournals.org
there is no intrinsic muscle mitochondrial defect in type 2
diabetic patients (7). On the other hand, high (postpran-
dial) levels of insulin increased the mitochondrial ATP
production rate in nondiabetic subjects, whereas this
increase was absent in type 2 diabetic patients (7). Fur-
thermore, the lack of response in the diabetic patients was
accompanied by a reduced expression of PGC-1, CS, and
COX.
Besides reduced insulin action, the hyperglycemia asso-
ciated with insulin resistance and type 2 diabetes has also
been suggested to exert harmful effects on mitochondrial
functional capacity via induction of oxidative stress. In-
deed, hyperglycemia has been shown to increase mito-
chondrial ROS production in endothelial cells (28), as well
as in other cell types (29). In addition, it was reported that
severe hyperglycemia inhibited respiration in human skel-
etal muscle, which was restored upon insulin treatment
(13).
It should be noted that the insulin-resistant state after
prolonged fasting was accompanied by hypoinsulinemia
and hypoglycemia. Mitochondrial function was thus
assessed after exposure to low insulin and glucose
concentrations. It is therefore unlikely that the reduced
mitochondrial functional capacity observed here is
caused by hyperinsulinemia and/or hyperglycemia asso-
ciated with reduced insulin action. However, it remains
possible that chronic hyperinsulinemia and hyperglyce-
mia may negatively affect mitochondrial function in type
2 diabetes patients.
An alternative candidate to explain the observed reduc-
tion in mitochondrial capacity upon fasting is prolonged
exposure to elevated plasma FFA levels. This is under-
scored by previous findings in isolated mouse and human
skeletal muscle mitochondria showing a dosage-depen-
dent inhibition of ATP synthesis upon incubation with high
but physiologic levels of FFA metabolites (30). Further-
more, it was shown in mice that prolonged consumption of
a high-fat diet for the duration of 16 weeks reduced
mitochondrial function (31). Also in human in vivo studies,
negative associations between high fatty acid availability
and markers for mitochondrial function have been re-
ported. Thus, PGC-1expression (9) and the insulin-
stimulated increase in skeletal muscle ATP synthesis (10)
were reduced upon lipid infusion. Furthermore, it was
recently shown that mitochondrial membrane potential
was impaired upon short-term lipid infusion in healthy
individuals, although several other markers of mitochon-
drial function remained unaffected (32). Despite the re-
ported negative associations between mitochondrial
function and (plasma) FFA, there are also several lines of
evidence suggesting the opposite. Thus, raising plasma
FFA by high-fat feeding combined with daily heparin
injections for 4 weeks in rats increased skeletal muscle
mitochondrial biogenesis and mitochondrial enzymes in-
volved in fat oxidation, the citric acid cycle, and the
respiratory chain (33). Furthermore, we previously
showed that high-fat feeding for 8 weeks in rats resulted in
a twofold increase of PGC-1protein levels (34).
Despite the obvious species differences between these
studies, the explanation for the discrepancy in these
results remains unclear. Adding to the complexity is the
fact that several approaches (high-fat feeding and lipid
infusion combined with a hyperinsulinemic-euglycemic
clamp) to elevate plasma FFA levels are accompanied by
hyperinsulinemia and/or hyperglycemia and insulin resis-
tance, all factors that have also been suggested to interfere
with mitochondrial capacity (26,29). Finally, differences in
absolute levels of plasma FFA achieved in the different
studies may contribute to (part of) the variation.
Within the context of mitochondrial lipotoxicity, we and
others have previously postulated that mitochondrial
UCP3 may be involved in protecting mitochondria against
(lipid-induced) oxidative damage (35). Therefore, we de-
termined protein levels of UCP3 and found that UCP3
content was similar between the fed and the fasted con-
dition. This is a surprising finding since fasting has been
quite convincingly shown to increase UCP3 protein levels
in animal studies (36,37). Moreover, UCP3 mRNA levels
were also elevated after 15 h (5-fold) and 40 h (10-fold)
of fasting in humans (38). It should be noted, however, that
this is the first study to evaluate UCP3 protein content
upon prolonged fasting in humans.
The impressive increase in plasma FFA upon pro-
longed fasting is in line with previous findings in hu-
mans (12,39), although the absolute values achieved in
this study (2.0 mol/l) are high. Also, the 2.7-fold
increase in IMTG levels after 60 h of fasting is slightly
higher in comparison with previous reports (12,39). The
high plasma FFA and IMTG levels might be caused by
complete compliance to the fasting regimen in the present
study since, in contrast to other studies, the subjects
stayed in a respiration chamber throughout the whole
period.
The reduction in insulin-stimulated glucose uptake ob-
served in the present study confirms previous observa-
tions showing that prolonged fasting reduced glucose Rd,
which was accounted for by a reduction in both insulin-
stimulated glucose oxidation and nonoxidative glucose
disposal (40). In agreement with previous reports (40), we
also detected a decreased metabolic flexibility (i.e., the
ability to switch from predominantly fat oxidation to
glucose oxidation upon insulin stimulation) upon pro-
longed fasting (Fig. 3). However, not all studies show this
effect (41).
As anticipated, whole-body fat oxidation increased sig-
nificantly upon prolonged fasting. Therefore, the decrease
in mitochondrial capacity in skeletal muscle is counterin-
tuitive, especially since this decrease was substrate-inde-
pendent and also apparent upon a lipid substrate.
These results indicate that the reduced mitochondrial
capacity is secondary to the fatty acid surplus associated
with the insulin-resistant state. It is important to note
however, that the reduction in muscle mitochondrial ca-
pacity does not (yet) affect the capability of the body to
enhance fat oxidation. This is an important finding since it
has generally been assumed that a reduction in muscle
mitochondrial function will result in reductions in whole-
body fat oxidative capacity (3–5,42,43). Here we show that
this extrapolation may not be justified, although we cannot
exclude that the fasting-induced reduced mitochondrial
function in muscle may decrease muscle-specific fat oxi-
dation compensated by increased fat oxidation in other
organs, or may have an impact on the capacity to switch
from carbohydrate to fat oxidation (metabolic flexibility).
In conclusion, 60 h of fasting in humans lowered insulin-
stimulated glucose uptake down to 50% along with drasti-
cally elevated plasma FFA and IMTG levels. This was
accompanied by an overall reduction in intrinsic mitochon-
drial functional capacity in skeletal muscle, despite a pro-
nounced increase in whole-body fat oxidation. Since
prolonged fasting is a physiologic condition in which in-
creased fat oxidation becomes very important, a reduced
J. HOEKS AND ASSOCIATES
diabetes.diabetesjournals.org DIABETES, VOL. 59, SEPTEMBER 2010 2123
mitochondrial function seems unbeneficial from a physio-
logic point of view. Our findings suggest that the elevated
plasma FFA and/or intramuscular lipid levels associated with
the insulin-resistant state are responsible for the secondary
negative effects on mitochondrial function.
ACKNOWLEDGMENTS
This study was funded by Top Institute Food and Nutri-
tion. J.H. was supported by a grant from the Netherlands
Organization for Scientific Research. M.K.C.H. is sup-
ported by a VIDI Research Grant for innovative research
from the Netherlands Organization for Scientific Research
(Grant 917.66.359). P.S. is supported by a VICI Research
Grant for innovative research from the Netherlands Orga-
nization for Scientific Research (Grant 918.96.618).
No potential conflicts of interest relevant to this article
were reported.
J.H. wrote the manuscript and researched data. N.A.v.H.
researched data and reviewed/edited the manuscript. M.M.
contributed to discussion and reviewed/edited the manu-
script. E.M.-K. and D.v.B. were involved in analysis of
material. M.K.C.H. reviewed/edited the manuscript. P.S.
edited the manuscript and contributed to discussion.
The authors thank Esther Phielix and Gert Schaart for
practical help and technical support.
REFERENCES
1. Petersen KF, Befroy D, Dufour S, Dziura J, Ariyan C, Rothman DL, DiPietro
L, Cline GW, Shulman GI. Mitochondrial dysfunction in the elderly:
possible role in insulin resistance. Science 2003;300:1140 –1142
2. Petersen KF, Dufour S, Befroy D, Garcia R, Shulman GI. Impaired
mitochondrial activity in the insulin-resistant offspring of patients with
type 2 diabetes. N Engl J Med 2004;350:664 671
3. Kelley DE, He J, Menshikova EV, Ritov VB. Dysfunction of mitochondria in
human skeletal muscle in type 2 diabetes. Diabetes 2002;51:2944 –2950
4. Phielix E, Schrauwen-Hinderling VB, Mensink M, Lenaers E, Meex R,
Hoeks J, Kooi ME, Moonen-Kornips E, Sels JP, Hesselink MK, Schrauwen
P. Lower intrinsic ADP-stimulated mitochondrial respiration underlies in
vivo mitochondrial dysfunction in muscle of male type 2 diabetic patients.
Diabetes 2008;57:2943–2949
5. Mogensen M, Sahlin K, Fernstrom M, Glintborg D, Vind BF, Beck-Nielsen
H, Hojlund K. Mitochondrial respiration is decreased in skeletal muscle of
patients with type 2 diabetes. Diabetes 2007;56:1592–1599
6. Pagel-Langenickel I, Bao J, Joseph JJ, Schwartz DR, Mantell BS, Xu X,
Raghavachari N, Sack MN. PGC-1alpha integrates insulin signaling, mito-
chondrial regulation, and bioenergetic function in skeletal muscle. J Biol
Chem 2008;283:22464 –22472
7. Asmann YW, Stump CS, Short KR, Coenen-Schimke JM, Guo Z, Bigelow
ML, Nair KS. Skeletal muscle mitochondrial functions, mitochondrial DNA
copy numbers, and gene transcript profiles in type 2 diabetic and nondia-
betic subjects at equal levels of low or high insulin and euglycemia.
Diabetes 2006;55:3309 –3319
8. Szendroedi J, Schmid AI, Chmelik M, Toth C, Brehm A, Krssak M, Nowotny
P, Wolzt M, Waldhausl W, Roden M. Muscle mitochondrial ATP synthesis
and glucose transport/phosphorylation in type 2 diabetes. PLoS Med
2007;4:e154
9. Hoeks J, Hesselink MK, Russell AP, Mensink M, Saris WH, Mensink RP,
Schrauwen P. Peroxisome proliferator-activated receptor-gamma coacti-
vator-1 and insulin resistance: acute effect of fatty acids. Diabetologia
2006;49:2419 –2426
10. Brehm A, Krssak M, Schmid AI, Nowotny P, Waldhausl W, Roden M.
Increased lipid availability impairs insulin-stimulated ATP synthesis in
human skeletal muscle. Diabetes 2006;55:136 –140
11. Johnson NA, Stannard SR, Mehalski K, Trenell MI, Sachinwalla T, Thomp-
son CH, Thompson MW. Intramyocellular triacylglycerol in prolonged
cycling with high- and low-carbohydrate availability. J Appl Physiol
2003;94:1365–1372
12. Stannard SR, Thompson MW, Fairbairn K, Huard B, Sachinwalla T,
Thompson CH. Fasting for 72 h increases intramyocellular lipid content in
nondiabetic, physically fit men. Am J Physiol Endocrinol Metab 2002;283:
E1185–1191
13. Rabol R, Hojberg PM, Almdal T, Boushel R, Haugaard SB, Madsbad S, Dela
F. Effect of hyperglycemia on mitochondrial respiration in type 2 diabetes.
J Clin Endocrinol Metab 2009;94:1372–1378
14. Hoeks J, van Baak MA, Hesselink MK, Hul GB, Vidal H, Saris WH,
Schrauwen P. Effect of beta1- and beta2-adrenergic stimulation on energy
expenditure, substrate oxidation, and UCP3 expression in humans. Am J
Physiol Endocrinol Metab 2003;285:E775–782
15. Kuipers H, Verstappen FT, Keizer HA, Geurten P, van Kranenburg G.
Variability of aerobic performance in the laboratory and its physiologic
correlates. Int J Sports Med 1985;6:197–201
16. Schoffelen PF, Westerterp KR, Saris WH, Ten Hoor F. A dual-respiration
chamber system with automated calibration. J Appl Physiol 1997;83:2064
2072
17. Schrauwen P, van Marken Lichtenbelt WD, Saris WH, Westerterp KR.
Changes in fat oxidation in response to a high-fat diet. Am J Clin Nutr
1997;66:276 –282
18. Bergstrom J, Hermansen L, Hultman E, Saltin B. Diet, muscle glycogen and
physical performance. Acta Physiol Scand 1967;71:140 –150
19. Frayn KN. Calculation of substrate oxidation rates in vivo from gaseous
exchange. J Appl Physiol 1983;55:628 634
20. Roorda BD, Hesselink MK, Schaart G, Moonen-Kornips E, Martinez-
Martinez P, Losen M, De Baets MH, Mensink RP, Schrauwen P. DGAT1
overexpression in muscle by in vivo DNA electroporation increases
intramyocellular lipid content. J Lipid Res 2005;46:230 –236
21. Koopman R, Schaart G, Hesselink MK. Optimisation of oil red O staining
permits combination with immunofluorescence and automated quantifica-
tion of lipids. Histochem Cell Biol 2001;116:63– 68
22. Shepherd D, PB G: Citrate synthase from rat liver. In Methods in
Enzymology. Vol XIII New York, Academic Press, 1969, p. 11–16
23. Boushel R, Gnaiger E, Schjerling P, Skovbro M, Kraunsoe R, Dela F.
Patients with type 2 diabetes have normal mitochondrial function in
skeletal muscle. Diabetologia 2007;50:790 –796
24. Schrauwen P, Mensink M, Schaart G, Moonen-Kornips E, Sels JP, Blaak
EE, Russell AP, Hesselink MK. Reduced skeletal muscle uncoupling
protein-3 content in prediabetic subjects and type 2 diabetic patients:
restoration by rosiglitazone treatment. J Clin Endocrinol Metab 2006;91:
1520 –1525
25. Bergman RN, Finegood DT, Ader M. Assessment of insulin sensitivity in
vivo. Endocr Rev 1985;6:45– 86
26. Stump CS, Short KR, Bigelow ML, Schimke JM, Nair KS. Effect of insulin
on human skeletal muscle mitochondrial ATP production, protein synthe-
sis, and mRNA transcripts. Proc Natl Acad SciUSA2003;100:7996 – 8001
27. Al-Khalili L, Forsgren M, Kannisto K, Zierath JR, Lonnqvist F, Krook A.
Enhanced insulin-stimulated glycogen synthesis in response to insulin,
metformin or rosiglitazone is associated with increased mRNA expression
of GLUT4 and peroxisomal proliferator activator receptor gamma co-
activator 1. Diabetologia 2005;48:1173–1179
28. Brownlee M. The pathobiology of diabetic complications: a unifying
mechanism. Diabetes 2005;54:1615–1625
29. Yu T, Robotham JL, Yoon Y. Increased production of reactive oxygen
species in hyperglycemic conditions requires dynamic change of mito-
chondrial morphology. Proc Natl Acad SciUSA2006;103:2653–2658
30. Abdul-Ghani MA, Muller FL, Liu Y, Chavez AO, Balas B, Zuo P, Chang Z,
Tripathy D, Jani R, Molina-Carrion M, Monroy A, Folli F, Van Remmen H,
DeFronzo RA. Deleterious action of FA metabolites on ATP synthesis:
possible link between lipotoxicity, mitochondrial dysfunction, and insulin
resistance. Am J Physiol Endocrinol Metab 2008;295:E678 685
31. Bonnard C, Durand A, Peyrol S, Chanseaume E, Chauvin MA, Morio B,
Vidal H, Rieusset J. Mitochondrial dysfunction results from oxidative
stress in the skeletal muscle of diet-induced insulin-resistant mice. J Clin
Invest 2008;118:789 – 800
32. Chavez AO, Kamath S, Jani R, Sharma LK, Monroy A, Abdul-Ghani MA,
Centonze VE, Sathyanarayana P, Coletta DK, Jenkinson CP, Bai Y, Folli F,
Defronzo RA, Tripathy D. Effect of short-term free fatty acids elevation on
mitochondrial function in skeletal muscle of healthy individuals. J Clin
Endocrinol Metab 2009
33. Garcia-Roves P, Huss JM, Han DH, Hancock CR, Iglesias-Gutierrez E, Chen
M, Holloszy JO. Raising plasma fatty acid concentration induces increased
biogenesis of mitochondria in skeletal muscle. Proc Natl Acad SciUSA
2007;104:10709 –10713
34. Hoeks J, Briede JJ, de Vogel J, Schaart G, Nabben M, Moonen-Kornips E,
Hesselink MK, Schrauwen P. Mitochondrial function, content and ROS
production in rat skeletal muscle: effect of high-fat feeding. FEBS Lett
2008;582:510 –516
35. Schrauwen P, Hoeks J, Hesselink MK. Putative function and physiological
relevance of the mitochondrial uncoupling protein-3: involvement in fatty
acid metabolism? Prog Lipid Res 2006;45:17– 41
PROLONGED FASTING AND MITOCHONDRIAL FUNCTION
2124 DIABETES, VOL. 59, SEPTEMBER 2010 diabetes.diabetesjournals.org
36. Moreno M, Lombardi A, De Lange P, Silvestri E, Ragni M, Lanni A, Goglia
F. Fasting, lipid metabolism, and triiodothyronine in rat gastrocnemius
muscle: interrelated roles of uncoupling protein 3, mitochondrial thioes-
terase, and coenzyme Q. FASEB J 2003;17:1112–1114
37. Cadenas S, Buckingham JA, Samec S, Seydoux J, Din N, Dulloo AG, Brand
MD. UCP2 and UCP3 rise in starved rat skeletal muscle but mitochondrial
proton conductance is unchanged. FEBS Lett 1999;462:257–260
38. Tunstall RJ, Mehan KA, Hargreaves M, Spriet LL, Cameron-Smith D.
Fasting activates the gene expression of UCP3 independent of genes
necessary for lipid transport and oxidation in skeletal muscle. Biochem
Biophys Res Commun 2002;294:301–308
39. Johnson NA, Stannard SR, Rowlands DS, Chapman PG, Thompson CH,
O’Connor H, Sachinwalla T, Thompson MW. Effect of short-term starvation
versus high-fat diet on intramyocellular triglyceride accumulation and
insulin resistance in physically fit men. Exp Physiol 2006;91:693–703
40. Bergman BC, Cornier MA, Horton TJ, Bessesen DH. Effects of fasting on
insulin action and glucose kinetics in lean and obese men and women.
Am J Physiol Endocrinol Metab 2007;293:E1103–1111
41. Soeters MR, Sauerwein HP, Dubbelhuis PF, Groener JE, Ackermans
MT, Fliers E, Aerts JM, Serlie MJ. Muscle adaptation to short-term
fasting in healthy lean humans. J Clin Endocrinol Metab 2008;93:2900
2903
42. Ortenblad N, Mogensen M, Petersen I, Hojlund K, Levin K, Sahlin K,
Beck-Nielsen H, Gaster M. Reduced insulin-mediated citrate synthase
activity in cultured skeletal muscle cells from patients with type 2 diabetes:
evidence for an intrinsic oxidative enzyme defect. Biochim Biophys Acta
2005;1741:206 –214
43. Ritov VB, Menshikova EV, He J, Ferrell RE, Goodpaster BH, Kelley DE.
Deficiency of subsarcolemmal mitochondria in obesity and type 2 diabetes.
Diabetes 2005;54:8 –14
J. HOEKS AND ASSOCIATES
diabetes.diabetesjournals.org DIABETES, VOL. 59, SEPTEMBER 2010 2125
... These findings agree with the data presented in Tables 1 and 3, suggesting that neither acute nor repeated fasting induce AMPK-PGC-1αmediated mitochondrial adaptations. Interestingly, fasting of 60-and 75-hours reduces complex I and II supported mitochondrial respiration in human skeletal muscle 50,51 , an effect that directly conflicts with the prevailing notion that fasting improves mitochondrial function 13 . Thus, fasting-mediated improvements in metabolic health in humans appear to be primarily mediated by caloric restriction 28 , and should notat least based on the evidence currently availablebe attributed to the induction of mitochondrial biogenesis or improved skeletal muscle mitochondrial respiratory capacity. ...
... In permeabilized muscle fibres, oxygen consumption was measured by high-resolution respirometry using an Oxygraph (Oroboros Instruments). A multisubstrate-uncoupler protocol with malate, octanoylcarnitine, ADP, glutamate, succinate and carbonylcyanide p-trifluoromethoxyphenylhydrazone (FCCP) was performed, as described previously (Hoeks et al., 2010). The integrity of the outer mitochondrial membrane was assessed as a quality control by the addition of cytochrome c upon maximal coupled respiration. ...
Article
Full-text available
Twenty‐four hour rhythmicity in whole‐body substrate metabolism, skeletal muscle clock gene expression and mitochondrial respiration is compromised upon insulin resistance. With exercise training known to ameliorate insulin resistance, our objective was to test if exercise training can reinforce diurnal variation in whole‐body and skeletal muscle metabolism in men with insulin resistance. In a single‐arm longitudinal design, 10 overweight and obese men with insulin resistance performed 12 weeks of high‐intensity interval training recurrently in the afternoon (between 14.00 and 18.00 h) and were tested pre‐ and post‐exercise training, while staying in a metabolic research unit for 2 days under free‐living conditions with regular meals. On the second days, indirect calorimetry was performed at 08.00, 13.00, 18.00, 23.00 and 04.00 h, muscle biopsies were taken from the vastus lateralis at 08.30, 13.30 and 23.30 h, and blood was drawn at least bi‐hourly over 24 h. Participants did not lose body weight over 12 weeks, but improved body composition and exercise capacity. Exercise training resulted in reduced 24‐h plasma glucose levels, but did not modify free fatty acid and triacylglycerol levels. Diurnal variation of muscle clock gene expression was modified by exercise training with period genes showing an interaction (time × exercise) effect and reduced mRNA levels at 13.00 h. Exercise training increased mitochondrial respiration without inducing diurnal variation. Twenty‐four‐hour substrate metabolism and energy expenditure remained unchanged. Future studies should investigate alternative exercise strategies or types of interventions (e.g. diet or drugs aiming at improving insulin sensitivity) for their capacity to reinforce diurnal variation in substrate metabolism and mitochondrial respiration. image Key points Insulin resistance is associated with blunted 24‐h flexibility in whole‐body substrate metabolism and skeletal muscle mitochondrial respiration, and disruptions in the skeletal muscle molecular circadian clock. We hypothesized that exercise training modifies 24‐h rhythmicity in whole‐body substrate metabolism and diurnal variation in skeletal muscle molecular clock and mitochondrial respiration in men with insulin resistance. We found that metabolic inflexibility over 24 h persisted after exercise training, whereas mitochondrial respiration increased independent of time of day. Gene expression of Per1 – 3 and Rorα in skeletal muscle changed particularly close to the time of day at which exercise training was performed. These results provide the rationale to further investigate the differential metabolic impact of differently timed exercise to treat metabolic defects of insulin resistance that manifest at a particular time of day.
... NF-κB activation, induced by in ammatory cytokines, can interfere with insulin signaling, promoting IR in muscles. 92 Mechanisms of muscular IR involve: mitochondrial dysfunction, ER stress, and serine phosphorylation IRS. Alterations in mitochondrial function and impaired oxidative capacity contribute to IR. Mitochondrial reactive oxygen species (ROS) production, disrupted lipid metabolism, and impaired mitochondrial biogenesis contribute to mitochondrial dysfunction-associated IR. 93 Perturbations in ER homeostasis and the UPR can trigger IR. ...
Preprint
Full-text available
Insulin resistance (IR) is a biological response to insulin stimulation in target tissues. IR alters glucose metabolism, resulting in increased insulin production by beta-cells. The primary condition associated with IR is obesity, which is often caused by environmental factors, particularly diet. Objective: To describe IR in various organs and present a signaling pathway project. Methods: The PubMed database was used to search for IR review publications. The referenced data for the signaling pathway were selected by aggregating references from the Kyoto Encyclopedia of Genes and Genomes (KEGG) database. A signaling pathway was designed based on IR research manuscripts, which show various mechanisms involved. The KEGG server was used to explore protein-protein interactions and create a signaling pathway diagram. The signaling path was mapped using PathVisio software, adapted to the model of the KEGG PATHWAY Database: https://www.genome.jp/pathway/map04930. Results: Articles featuring the terms “insulin resistance” and “signaling pathway” were selected from the PubMed database. Based on validated research articles, well-founded pathways were chosen and a representative description of these pathways was achieved. Reproduction contigs from the KEGG database projected the signaling pathway of biomolecules leading to IR. Thus, the interaction between multiple mechanisms releases factors that contribute to the development of IR. Conclusion: The interaction between multiple mechanisms and molecular interactions are important factors in the development of IR in various organs and systems.
... Permeabilized muscle fibers were prepared from the muscle tissue collected in the BIOPS preservation medium, as described previously (van de Weijer et al., 2015). Subsequently, the permeabilized muscle fibers (~2.5 mg wet weight) were analyzed for ex vivo mitochondrial respiration assessment using high-resolution respirometry (Oxygraph, OROBOROS Instruments; Hoeks et al., 2010). To prevent oxygen deprivation during the measurement, the respiration chambers were hyperoxygenated up to ~400 μmol/L O 2 . ...
Article
Full-text available
Mitochondria are organelles that fuel cellular energy requirements by ATP formation via aerobic metabolism. Given the wide variety of methods to assess skeletal muscle mitochondrial capacity, we tested how well different invasive and noninvasive markers of skeletal muscle mitochondrial capacity reflect mitochondrial respiration in permeabilized muscle fibers. Nineteen young men (mean age: 24 ± 4 years) were recruited, and a muscle biopsy was collected to determine mitochondrial respiration from permeabilized muscle fibers and to quantify markers of mitochondrial capacity, content such as citrate synthase (CS) activity, mitochondrial DNA copy number, TOMM20, VDAC, and protein content for complex I-V of the oxidative phosphorylation (OXPHOS) system. Additionally, all participants underwent noninvasive assessments of mitochondrial capacity: PCr recovery postexercise (by 31 P-MRS), maximal aerobic capacity, and gross exercise efficiency by cycling exercise. From the invasive markers, Complex V protein content and CS activity showed the strongest concordance (Rc = 0.50 to 0.72) with ADP-stimulated coupled mitochondrial respiration, fueled by various substrates. Complex V protein content showed the strongest concordance (Rc = 0.72) with maximally uncoupled mitochondrial respiration. From the noninvasive markers, gross exercise efficiency, VO2max , and PCr recovery exhibited concordance values between 0.50 and 0.77 with ADP-stimulated coupled mitochondrial respiration. Gross exercise efficiency showed the strongest concordance with maximally uncoupled mitochondrial respiration (Rc = 0.67). From the invasive markers, Complex V protein content and CS activity are surrogates that best reflect skeletal muscle mitochondrial respiratory capacity. From the noninvasive markers, exercise efficiency and PCr recovery postexercise most closely reflect skeletal muscle mitochondrial respiratory capacity.
Article
Full-text available
Dysregulated metabolic dynamics are evident in both cancer and diabetes, with metabolic alterations representing a facet of the myriad changes observed in these conditions. This review delves into the commonalities in metabolism between cancer and type 2 diabetes (T2D), focusing specifically on the contrasting roles of oxidative phosphorylation (OXPHOS) and glycolysis as primary energy-generating pathways within cells. Building on earlier research, we explore how a shift towards one pathway over the other serves as a foundational aspect in the development of cancer and T2D. Unlike previous reviews, we posit that this shift may occur in seemingly opposing yet complementary directions, akin to the Yin and Yang concept. These metabolic fluctuations reveal an intricate network of underlying defective signaling pathways, orchestrating the pathogenesis and progression of each disease. The Warburg phenomenon, characterized by the prevalence of aerobic glycolysis over minimal to no OXPHOS, emerges as the predominant metabolic phenotype in cancer. Conversely, in T2D, the prevailing metabolic paradigm has traditionally been perceived in terms of discrete irregularities rather than an OXPHOS-to-glycolysis shift. Throughout T2D pathogenesis, OXPHOS remains consistently heightened due to chronic hyperglycemia or hyperinsulinemia. In advanced insulin resistance and T2D, the metabolic landscape becomes more complex, featuring differential tissue-specific alterations that affect OXPHOS. Recent findings suggest that addressing the metabolic imbalance in both cancer and diabetes could offer an effective treatment strategy. Numerous pharmaceutical and nutritional modalities exhibiting therapeutic effects in both conditions ultimately modulate the OXPHOS–glycolysis axis. Noteworthy nutritional adjuncts, such as alpha-lipoic acid, flavonoids, and glutamine, demonstrate the ability to reprogram metabolism, exerting anti-tumor and anti-diabetic effects. Similarly, pharmacological agents like metformin exhibit therapeutic efficacy in both T2D and cancer. This review discusses the molecular mechanisms underlying these metabolic shifts and explores promising therapeutic strategies aimed at reversing the metabolic imbalance in both disease scenarios.
Article
Type 2 diabetes mellitus (DM) is a common chronic condition characterized by persistent hyperglycemia and is associated with insulin resistance (IR) in critical glucose-consuming tissues, including skeletal muscle and adipose tissue. Oxidative stress and mitochondrial dysfunction are known to play key roles in IR. Acrolein is a reactive aldehyde found in the diet and environment that is generated as a fatty acid product through the glucose autooxidation process under hyperglycemic conditions. Our previous studies have shown that acrolein impairs insulin sensitivity in normal and diabetic mice, and this effect can be reversed by scavenging acrolein. This study demonstrated that acrolein increased oxidative stress and inhibited mitochondrial respiration in differentiated C2C12 myotubes and differentiated 3T3-L1 adipocytes. As a result, insulin signaling pathways were inhibited, leading to reduced glucose uptake. Treatment with acrolein scavengers, N-acetylcysteine, or carnosine ameliorated mitochondrial dysfunction and inhibited insulin signaling. Additionally, an increase in acrolein expression correlated with mitochondrial dysfunction in the muscle and adipose tissues of diabetic mice. These findings suggest that acrolein-induced mitochondrial dysfunction contributes to IR, and scavenging acrolein is a potential therapeutic approach for treating IR.
Chapter
Skeletal muscle biopsy is a long-established diagnostic used primarily as a diagnostic tool for neuromuscular disorders characterized by reduced muscle function and strength. For anatomical and functional characteristics, leg muscles and especially the vastus lateralis have been most commonly investigated. Percutaneous needles, which overcame the more invasive open biopsies, were introduced more than 50 years ago, with the original instruments named after Bergstrom [1], in honor of his pioneering work in their development (Fig. 1). When adequate suction is applied and a sufficient amount of muscle tissue is recovered, muscle biopsy allows for multiple measurements as well as the assessment of different anatomical or physiological parameters. Fiber and cell isolation, incubation, or culture are also possible and enable additional ex vivo studies. Muscular dystrophies, mitochondrial myopathies, and conditions often characterized by impaired muscle strength and function were early and obvious targets for diagnostic and research applications of muscle biopsy. Needle biopsy has been further extensively applied in the study of exercise physiology and pathophysiology, with the goal of investigating the regulation of mitochondrial function and substrate oxidation. In recent years, studies in the fields of obesity and diabetes have also focused on muscle mitochondrial function, and muscle biopsies have become increasingly common in human metabolic assessments. This chapter provides an overview of mitochondrial function and substrate utilization while addressing the major concepts emerging from basic studies on mitochondrial regulation. Available data from human skeletal muscle biopsy studies are reviewed, with particular emphasis placed on the effects of nutritional state, diet, and exercise and their potential interactions with insulin resistance and disease state.KeywordsSkeletal muscle biopsySkeletal muscle functionMitochondriaMitochondrial glucoseFatty acid oxidationMitochondrial oxidative metabolism
Article
Objectives: The aim of this study is to investigate the effect of intermittent fasting (IF) on the regulation of skeletal muscle protein metabolism in response to nutrient supplementation during fasting. Methods: Twelve-week-old male C57BL/6J mice were assigned to two groups: ad libitum and IF, with the latter having access to food for only 3 h/d. After 6 wk of experimental periods, an oral glucose tolerance test was performed. One week later, phosphate-buffered saline or a glucose and branched-chain amino acid mixture was administered orally, and blood and tissues were collected 30 min later. Results: The oral glucose tolerance test results revealed that the IF group had better insulin sensitivity. They also had lower body and fat weights while maintaining the same level of skeletal muscle mass as the ad libitum group. The phosphorylation of ribosomal protein S6 in the skeletal muscle, a marker for the activation of protein translation, was greater in the IF group after glucose and branched-chain amino acid mixture administration. Microtubule-associated protein light chain 3-II-to-light chain 3-I ratio, a marker for autophagosome formation, in skeletal muscle during fasting was significantly lower in the IF group than that in the ad libitum group. Conclusions: Our findings suggest that adaptation to IF regulates protein synthesis and breakdown, leading to the maintenance of skeletal muscle mass while reducing body fat.
Article
Full-text available
The pathophysiology underlying mitochondrial dysfunction in insulin-resistant skeletal muscle is incompletely characterized. To further delineate this we investigated the interaction between insulin signaling, mitochondrial regulation, and function in C2C12 myotubes and in skeletal muscle. In myotubes elevated insulin and glucose disrupt insulin signaling, mitochondrial biogenesis, and mitochondrial bioenergetics. The insulin-sensitizing thiazolidinedione pioglitazone restores these perturbations in parallel with induction of the mitochondrial biogenesis regulator PGC-1alpha. Overexpression of PGC-1alpha rescues insulin signaling and mitochondrial bioenergetics, and its silencing concordantly disrupts insulin signaling and mitochondrial bioenergetics. In primary skeletal myoblasts pioglitazone also up-regulates PGC-1alpha expression and restores the insulin-resistant mitochondrial bioenergetic profile. In parallel, pioglitazone up-regulates PGC-1alpha in db/db mouse skeletal muscle. Interestingly, the small interfering RNA knockdown of the insulin receptor in C2C12 myotubes down-regulates PGC-1alpha and attenuates mitochondrial bioenergetics. Concordantly, mitochondrial bioenergetics are blunted in insulin receptor knock-out mouse-derived skeletal myoblasts. Taken together these data demonstrate that elevated glucose and insulin impairs and pioglitazone restores skeletal myotube insulin signaling, mitochondrial regulation, and bioenergetics. Pioglitazone functions in part via the induction of PGC-1alpha. Moreover, PGC-1alpha is identified as a bidirectional regulatory link integrating insulin-signaling and mitochondrial homeostasis in skeletal muscle.
Article
Full-text available
Mitochondrial dysfunction has been proposed as an underlying mechanism in the pathogenesis of insulin resistance and type 2 diabetes mellitus. To determine whether mitochondrial dysfunction plays a role in the free fatty acid (FFA)-induced impairment in insulin action in skeletal muscle of healthy subjects. Eleven lean normal glucose tolerant individuals received 8 h lipid and saline infusion on separate days with a euglycemic insulin clamp during the last 2 h. Vastus lateralis muscle biopsies were performed at baseline and after 6 h lipid or saline infusion. Inner mitochondrial membrane potential (Psi(m)) and mitochondrial mass were determined ex vivo by confocal microscopy. Compared with saline infusion, lipid infusion reduced whole-body glucose uptake by 22% (P < 0.05). Psi(m) decreased by 33% (P < 0.005) after lipid infusion and the decrement in Psi(m) correlated with change in plasma FFA after lipid infusion (r = 0.753; P < 0.005). Mitochondrial content and morphology did not change after lipid infusion. No significant changes in genes expression, citrate synthase activity, and total ATP content were observed after either lipid or saline infusion. Short-term physiological increase in plasma FFA concentration in lean normal glucose tolerant subjects induces insulin resistance and impairs mitochondrial membrane potential but has no significant effects on mitochondrial content, gene expression, ATP content, or citrate synthase activity.
Chapter
This chapter discusses the citrate synthase from rat liver. Citrate synthase has been well characterized previously from a variety of sources, including pig heart, pig liver, pigeon breast muscle, moth flight muscle, yeast and Escherichia coli, and a regulatory function has been postulated as it is effected allosterically by adenine nucleotides or NADH. Evidence that citrate synthase controls the flow of acetyl-CoA in the tricarboxylic acid cycle is presented only in the case of rat liver, and the enzyme may not have a regulatory role in those tissues where its activity is in considerable excess (20- to 40-fold) of the maximal rates of acetyl-CoA production or isocitrate oxidation. Citrate synthase in rat liver is localized intramitochondrially, and hence the first step in the purification of the enzyme is the preparation of mitochondria, as this reduces the amount of starting protein by 50- 60% without any loss in activity.
Article
The muscle glycogen content of the quadriceps femoris muscle was determined in 9 healthy subjects with the aid of the needle biopsy technique. The glycogen content could be varied in the individual subjects by instituting different diets after exhaustion of the glycogen store by hard exercise. Thus, the glycogen content after a fat ± protein (P) and a carbohydrate-rich (C) diet varied maximally from 0.6 g/100g muscle to 4.7 g. In all subjects, the glycogen content after the C diet was higher than the normal range for muscle glycogen, determined after the mixed (M) diet. After each diet period, the subjects worked on a bicycle ergometer at a work load corresponding to 75 per cent of their maximal O2 uptake, to complete exhaustion. The average work time was 59, 126 and 189 min after diets P, M and C, and a good correlation was noted between work time and the initial muscle glycogen content. The total carbohydrate utilization during the work periods (54–798 g) was well correlated to the decrease in glycogen content. It is therefore concluded that the glycogen content of the working muscle is a determinant for the capacity to perform long-term heavy exercise. Moreover, it has been shown that the glycogen content and, consequently, the long-term work capacity can be appreciably varied by instituting different diets after glycogen depletion.
Article
Increased production of mitochondrial reactive oxygen species (ROS) by hyperglycemia is recognized as a major cause of the clinical complications associated with diabetes and obesity [Brownlee, M. (2001) Nature 414, 813–820]. We observed that dynamic changes in mitochondrial morphology are associated with high glucose-induced overproduction of ROS. Mitochondria undergo rapid fragmentation with a concomitant increase in ROS formation after exposure to high glucose concentrations. Neither ROS increase nor mitochondrial fragmentation was observed after incubation of cells with the nonmetabolizable stereoisomer l-glucose. However, inhibition of mitochondrial pyruvate uptake that blocked ROS increase did not prevent mitochondrial fragmentation in high glucose conditions. Importantly, we found that mitochondrial fragmentation mediated by the fission process is a necessary component for high glucose-induced respiration increase and ROS overproduction. Extended exposure to high glucose conditions, which may mimic untreated diabetic conditions, provoked a periodic and prolonged increase in ROS production concomitant with mitochondrial morphology change. Inhibition of mitochondrial fission prevented periodic fluctuation of ROS production during high glucose exposure. These results indicate that the dynamic change of mitochondrial morphology in high glucose conditions contributes to ROS overproduction and that mitochondrial fission/fusion machinery can be a previously unrecognized target to control acute and chronic production of ROS in hyperglycemia-associated disorders. • DLP1/Drp1 • mitochondrial fission • dynamin • diabetes • obesity