You are on page 1of 26

Article

Excessive exercise training causes mitochondrial


functional impairment and decreases glucose
tolerance in healthy volunteers
Graphical abstract Authors
Mikael Flockhart, Lina C. Nilsson,
Senna Tais, Björn Ekblom,
William Apró, Filip J. Larsen

Correspondence
mikael.flockhart@gih.se (M.F.),
filip.larsen@gih.se (F.J.L.)

In brief
Flockhart and colleagues investigated the
dose-response relationship of exercise
training and found that excessive training
causes mitochondrial impairment and
reduced glucose tolerance in healthy
volunteers. Mitochondrial ROS
production was reduced, thereby
maintaining total oxidative stress. In a
separate cohort, disturbed glucose
control was found in free-living elite
endurance athletes.

Highlights
d Here, we investigate the dose-response effects of exercise
training in healthy subjects

d Excessive exercise training induces substantial


mitochondrial respiratory impairment

d Mitochondrial impairment is associated with impaired


glucose tolerance

d Despite excessive training, markers of global oxidative stress


were unchanged

Flockhart et al., 2021, Cell Metabolism 33, 957–970


May 4, 2021 ª 2021 Elsevier Inc.
https://doi.org/10.1016/j.cmet.2021.02.017 ll
ll

Article
Excessive exercise training causes mitochondrial
functional impairment and decreases
glucose tolerance in healthy volunteers
Mikael Flockhart,1,* Lina C. Nilsson,1 Senna Tais,1 Björn Ekblom,1 William Apró,1,2 and Filip J. Larsen1,3,*
1The Swedish School of Sport and Health Sciences, GIH, Åstrand Laboratory, Department of Physiology, Nutrition and Biomechanics,
Stockholm 114 33, Sweden
2Department of Clinical Sciences, Intervention and Technology, Karolinska Institutet, Stockholm, Sweden
3Lead contact

*Correspondence: mikael.flockhart@gih.se (M.F.), filip.larsen@gih.se (F.J.L.)


https://doi.org/10.1016/j.cmet.2021.02.017

SUMMARY

Exercise training positively affects metabolic health through increased mitochondrial oxidative capacity and
improved glucose regulation and is the first line of treatment in several metabolic diseases. However, the up-
per limit of the amount of exercise associated with beneficial therapeutic effects has not been clearly iden-
tified. Here, we used a training model with a progressively increasing exercise load during an intervention
over 4 weeks. We closely followed changes in glucose tolerance, mitochondrial function and dynamics, phys-
ical exercise capacity, and whole-body metabolism. Following the week with the highest exercise load, we
found a striking reduction in intrinsic mitochondrial function that coincided with a disturbance in glucose
tolerance and insulin secretion. We also assessed continuous blood glucose profiles in world-class endur-
ance athletes and found that they had impaired glucose control compared with a matched control group.

INTRODUCTION The quality of the mitochondrial population and its effects on


metabolic health is not merely regulated by the mitochondrial
Mitochondria are the primary sources of cellular ATP synthesis number and its respiratory capacity. An emerging field of
and are major ROS (reactive oxygen species)-producing organ- research has shown that the location and function of the mito-
elles and are thus central in energy and redox metabolism. The chondria within the cell is highly dynamic and regulated by fusion
size and function of the mitochondrial pool are vital for metabolic and fission events, creating a large communicative mitochon-
health and muscular function. Mitochondrial capacity is tightly drial network (Westermann, 2010). In many pathological states,
correlated to whole-body maximal oxygen uptake (van der mitochondria display a reduced ability to respond to nutrient
Zwaard et al., 2016), which itself is a strong proxy for metabolic supply and match mitochondrial respiration to metabolic de-
function and health (Lee et al., 2011). A reduced mitochondrial mand (Nunnari and Suomalainen, 2012). In addition, the tran-
function is often (Bhatti et al., 2017; Kelley et al., 2002; Lowell scription factor NF-E2 p45-related factor 2 (Nrf2) and the E3
and Shulman, 2005; Simoneau et al., 1999), but not always ligase adaptor Kelch-like ECH-associated protein 1 (KEAP1),
(Boushel et al., 2007; Lefort et al., 2010), observed in insulin- which is the primary negative regulator of Nrf2, are both critical
resistant subjects. Studies using transgenic rodents with severe players in redox homeostasis, the antioxidative response, and
reductions in mitochondrial content have a paradoxically mitochondrial biogenesis. It was recently shown that acute exer-
improved glucose tolerance (Zechner et al., 2010). A mismatch cise increases the total Nrf2 protein content within minutes after
between fatty acid availability and mitochondrial fat oxidation an exercise bout (Gallego-Selles et al., 2020), but less is known
has been implicated in the accumulation of ectopic fat and cer- regarding the chronic response to training.
amide, which is a common finding in insulin-resistant skeletal Exercise training has proven to be a powerful tool to stimulate
muscles (Pan et al., 1997). Therefore, it is highly debated whether mitochondrial biogenesis and can act as a preventative treat-
mitochondrial dysfunction itself is a cause or a consequence of ment against many metabolic disorders by stimulating glucose
impaired blood glucose control and insulin sensitivity. The term uptake (Pedersen and Saltin, 2015). However, an upper limit
‘‘mitochondrial dysfunction’’ has no clear definition, and mito- where the exercise stimuli no longer result in further positive
chondrial efficiency and ROS production (Konopka et al., metabolic outcomes has not been clearly identified. The benefits
2015), intrinsic mitochondria respiration (Mogensen et al., of exercise training on cardiac health seem to follow a curvilinear
2007), and mitochondrial density (Morino et al., 2005) have all relationship where modest to high investments in physical activ-
been reported to differ between insulin-resistant subjects and ity appear to be most beneficial. High to extreme amounts of ex-
controls. ercise training have been associated with negative effects on

Cell Metabolism 33, 957–970, May 4, 2021 ª 2021 Elsevier Inc. 957
ll
Article

cardiac health that include increased coronary artery calcifica- With regard to adaptations on the whole-body level, our data
tion, myocardial fibrosis, and arrhythmia. Although a reversed show the expected physiological changes after a period with
J-shaped association between the amount of exercise and progressively harder training, commonly referred to as ‘‘over-
health outcomes have been demonstrated, no clear threshold reaching,’’ and after a week with drastically reduced training,
between optimal and excessive amounts of exercise has been the performance was increased in line with the concept of super-
established (Eijsvogels et al., 2018; Franklin et al., 2020). compensation (Issurin, 2010).
In competitive sports, it is well established that an appropriate
exercise stimulus improves performance, whereas too much Mitochondrial respiration is reduced after excessive
training leads to staleness and reduced performance. In the training with a compensatory increase in markers of
short-term perspective, this state is termed as nonfunctional mitochondrial density
overreaching, which can develop into overtraining syndrome in The classic cellular response to exercise training is an increase in
the long term (Halson and Jeukendrup, 2004). In a study in which the amount and respiratory capacity of mitochondria that in-
12 subjects performed sprint interval training, we found a sub- creases the respiratory capacity of a given unit of muscle (Hol-
stantial reduction in mitochondrial respiration with oxidative al- loszy, 1967). Intrinsic mitochondrial respiration (IMR) is instead
terations of mitochondrial proteins (Larsen et al., 2016). As the the respiratory capacity related to a mitochondrial parameter,
mitochondrial function is tightly associated with metabolic often a mitochondrial protein or citrate synthase (CS) activity.
health, such derangements should have negative metabolic out- We have previously shown an impaired IMR after intense
comes. Therefore, we hypothesized that there is a bell-shaped short-term sprint interval training (Larsen et al., 2016), and scat-
relationship between exercise training load and mitochondrial tered reports of impaired IMR have been found after extreme
function, glucose metabolism, and physiological adaptation to endurance events (Konopka et al., 2017) as well as acutely after
exercise training in human subjects, during a training program a 5-km cycling time trial (Layec et al., 2018). Knowing that phys-
with a progressive increase in training load. ical performance and VO2max increased in an expected manner
during the training period, we wanted to determine how these
changes were reflected at the mitochondrial level. We, therefore,
RESULTS
measured ex vivo IMR in isolated mitochondria from vastus later-
alis, to determine how this parameter responded to the different
To test our hypothesis, we recruited six female and five male
training phases in this study.
healthy subjects to take part in a 4-week training intervention
During the first part of the intervention, from BL to LT and MT,
consisting of high-intensity interval training (HIIT; see Table S1
IMR numerically increased (from 13.2 ± 1.8 to 15.7 ± 1.4
for subject characteristics). During the first 3 weeks, we carefully
pmol$s 1$mg 1) but decreased drastically by 40% (to 9.4 ± 1.1
titrated the exercise sessions more and more often and were
pmol$s 1$mg 1) during ET compared with MT (Figure 2A). In
thereby able to study the physiological responses to different
RE, intrinsic mitochondrial function seemed to recover numeri-
loads of exercise training. During the fourth week, the exercise
cally but was still 25% lower than that during MT (11.8 ± 1.1
load was reduced to allow for recovery. Throughout the interven-
pmol$s 1$mg 1) (Figure 2A). Compared with MT, the reduction
tion, muscle biopsies and oral glucose tolerance tests (OGTTs)
in mitochondrial intrinsic respiration after ET was evident,
were performed to assess the metabolic response in different
regardless of the substrate combination used, which decreased
phases. See Figure 1A for schematics of the study design, Fig-
by 45% (from 8.3 ± 0.9 to 4.6 ± 0.5 pmol$s 1$mg 1) when
ure 1B for exercise training load, and Figure S1 for the time
respiring on complex I substrates (Figure 2B) and decreased
points of biopsy and OGTTs. A detailed description of the
by 42% (from 9.0 ± 0.9 to 5.2 ± 0.4 pmol$s 1$mg 1) in the pres-
method is presented in the STAR methods section.
ence of succinate and rotenone that support complex II respira-
tion (Figure 2C). Similar reductions in IMR were seen when
Performance and cardiovascular adaptations to respiration was normalized to CS activity in the mitochondrial
progressive HIIT suspension (Figure S2). However, in LEAK state (in the absence
The HIIT used during the study was proven effective in enhancing of ADP but in the presence of respiratory substrates pyruvate,
performance and physiological capacities. Power output during glutamate, and malate) (Figure 2D) as well as for ADP-stimulated
HIIT increased from baseline (BL) through light training (LT) and respiration on the lipid substrate octanoyl carnitine and malate,
moderate training (MT) (Figures 1C and 1D). In the phase of no decreases were found in IMR after ET (Figure 2E). The respi-
excessive training (ET), despite increasing the training load, the ratory control index was 17.3 ± 1.1 on an average and was unaf-
improvement in physical performance stagnated, indicating fected throughout the intervention (Figure 2F), and the externally
what others have described as accumulating fatigue and malad- added cytochrome c had no significant effect on respiration (Fig-
aptation to the exercise stimuli (Halson and Jeukendrup, 2004). ure 2G). This indicates a high quality of our mitochondrial fraction
After the recovery phase (RE), when the training load was dras- that remained consistent during the intervention.
tically reduced, performance peaked (Figure 1C). In contrast to The amount of protein and activity of the mitochondrial
the dynamic response in physical performance, maximal oxygen enzyme CS is often used as a quantitative estimate of the mito-
consumption increased consistently throughout the study, chondrial content and serves as a stable marker for long-term
regardless of the training phase (Figure 1E). Maximal attainable adaptations to the endurance exercise (Vigelsø et al., 2014). Us-
heart rate (Figure 1F) during HIIT training was suppressed after ing a spectrophotometric approach, we measured CS activity in
ET but was restored after the RE, and heart rate response to sub- muscle homogenate and observed a steady increase throughout
maximal work followed a similar pattern (Figure 1G). the training period (Figure 2H). To further assess the effect of

958 Cell Metabolism 33, 957–970, May 4, 2021


ll
Article

Figure 1. Schematics of study design and performance and cardiovascular adaptations to progressive HIIT
(A) During the intervention, 14 sessions of HIIT training were performed on a cycle ergometer, of which 6 were combined with power output and physiological
measurements of VO2, CO2, blood lactate, blood glucose, and heart rate. HIIT sessions started with a standardized warm-up of 10 min at 100 W at 70 rpm,
followed by five 4-min-long intervals at ~95% of VO2max with 3 min of rest. The HIIT sessions were scheduled in order to capture mitochondrial status and
physiological outcomes at BL, after an initial LT load, after MT load, after ET load, and after a period of RE in training load. During the training intervention, five
biopsies from vastus lateralis were taken and four OGTTs were conducted. Each biopsy and OGTT was performed in fasted state 14 h post exercise.
(B) The density of HIIT in the different phases expressed as minutes of HIIT per week.
(C and D) Cycling performance measured as mean power output during HIIT sessions (C) and change in power output session to session (D).
(E) Mean oxygen uptake during HIIT sessions.
(F) Maximal heart rate during HIIT sessions.
(G) Heart rate at a fixed submaximal intensity of 100 W.
Unless otherwise stated, a repeated-measure one-way ANOVA was used, and a significant main effect (of time) is marked with a # in graphs. Dunnett’s post hoc
test was used on normally distributed data and Dunn’s post hoc test was used on non-normal distributed data. Post hoc tests were performed between ET and all
other situations. All values are mean ± SEM. For all measurements, n = 11.

Cell Metabolism 33, 957–970, May 4, 2021 959


ll
Article

Figure 2. Mitochondrial respiration decreases after excessive training with a compensatory increase in markers of mitochondrial density
(A) Maximal ADP-stimulated respiration for complex I+II in the presence of the respiratory substrates pyruvate, glutamate, succinate, and malate.
(B) Respiration through complex I using pyruvate, glutamate, and malate.
(C) Respiration through complex II using succinate as a respiratory substrate in the presence of rotenone to inhibit complex I.
(D) LEAK respiration (in the presence of pyruvate, glutamate, and malate but without added ADP).
(E) Fat oxidation stimulated by octanoyl carnitine, malate, and ADP.
(F) The respiratory control indices were calculated as complex I+II respiration divided by leak respiration.
(G) Addition of cytochrome c to assess membrane integrity in the mitochondrial fraction.
(H) The activity of citrate synthase in muscle homogenate was assessed spectrophotometrically.
(I–L) The abundance of CS protein (I), the mitochondrial outer membrane channel protein VDAC1 (J), and inner membrane organizer Mitofilin (K) were all measured
by the western blot technique. A reference memcode staining is presented in (L). For all measurements, n = 11.

exercise training on the abundance of mitochondrial proteins, we pected manner and stands in sharp contrast with the observed
also performed western blot analysis of CS protein, the voltage- decrease in IMR after the ET phase.
dependent anion-selective channel 1 (VDAC1) located in the
outer mitochondrial membrane, mitochondrial inner membrane Excessive training reduces glucose tolerance while
organizer Mitofilin, and proteins in the electron transport chain. maintaining metabolic flexibility during a glucose load
The abundance of CS protein displayed a numerical but nonsig- The mitochondrial dysfunction that is observed in T2D patients is
nificant increase during the intervention (Figure 2I), whereas an intensely debated potential cause of insulin resistance in skel-
VDAC1, Mitofilin, and complex I–IV increased throughout the etal muscle (Pinti et al., 2019). If the mitochondrial function is an
intervention (Figures 2J, 2K, and S3). Conclusively, mitochon- important determinant of insulin sensitivity, the dramatic impair-
drial content and enzymatic activity increased stepwise in an ex- ment in IMR observed in this study could translate into a

960 Cell Metabolism 33, 957–970, May 4, 2021


ll
Article

Figure 3. Changes in glucose tolerance, insulin secretion, GLUT4 abundance, and metabolic flexibility following progressive exercise
training
(A and B) Glucose tolerance was tested by administering a 75 g glucose solution in water after an overnight fast. Glucose was measured in venous blood samples
and the glucose area under the curve was calculated during the following 120 min.
(C) Correlation analysis of the change in mitochondrial respiration and the change in AUC for glucose between the different situations. The baseline values are
represented by the dot in origin. Group mean values in each situation are presented in the graph.
(D) Insulin was measured using a colorimetric ELISA assay.
(E and F) HOMA-IR (E) and HOMA-b (F) was calculated from glucose and insulin values at each time point.
(G) C-peptide was measured at each time point during the OGTT.
(H and I) GLUT4 protein in whole-muscle homogenate (H) and in a membrane-bound fraction (I) was measured using western blot. Memcode staining was used as
a loading control.

(legend continued on next page)

Cell Metabolism 33, 957–970, May 4, 2021 961


ll
Article

disturbance in glucose control. Indeed, the 40% reduction in Insulin resistance and type 2 diabetes (T2D) have been re-
IMR observed in ET surpasses the difference of ~20% in IMR ported to be associated with a reduced capacity to oxidize lipid
between T2D patients and healthy weight-matched controls pre- substrates in some studies (Kelley and Simoneau, 1994; Men-
viously reported (Mogensen et al., 2007). To elucidate if the sink et al., 2001). Others report no difference between type 2 di-
reduction in IMR in this study was associated with a disturbed abetics and healthy subjects (Mogensen et al., 2007), whereas
metabolic function, we performed repeated OGTT in combina- some have found increased lipid oxidation and accumulation
tion with indirect calorimetry at four different time points (BL, of incomplete beta-oxidation metabolites in the insulin-resistant
LT, ET, and RE) throughout the training intervention. We found state (Koves et al., 2008). Since the capacity to oxidize lipids is
that glucose area under the curve (AUC) was unaltered during not maximally stimulated during resting conditions, we wanted
the first part of the intervention (from 600 ± 29 at BL to 589 ± to measure this capacity at a submaximal work rate (100 W).
22 mmol$L 1$min after LT) but were strikingly increased to This work rate corresponded to 48% ± 2% of the subjects
658 ± 23 mmol$L 1$min during ET (Figures 3A and 3B). Interest- VO2max and is close to the relative work rate that elicits the high-
ingly, during the RE phase, glucose tolerance only partly recov- est rates of lipid oxidation (Jeukendrup and Wallis, 2005). We
ered, closely resembling the dysfunction in mitochondrial found that the rate of lipid oxidation increased continuously
respiration that showed the same pattern. In a sub-analysis, from BL (5.6 ± 0.9 kJ$min 1) and was the highest after ET
we found that the change in glucose tolerance was significantly (11.0 ± 0.9 kJ$min 1), despite the substantial reduction in IMR
correlated to the change in mitochondrial respiration, time point at this time point (Figure 3L). These results indicate that the ca-
to time point (Figure 3C). pacity to oxidize lipids is not responsible for the reduced glucose
The AUC for plasma insulin changed during the intervention tolerance in this setting. The subjects remained weight stable
and decreased from 6,141 ± 890 after LT to 4,950 ± during the intervention (69.4 ± 11.2 kg at BL and 69.0 ±
556 mU$mL 1$min after ET (Figure 3D). Therefore, the calculated 11.0 kg at ET; Figure S6) and, therefore, we rule out the possibil-
HOMA-IR values were unchanged (Figure 3E) but HOMA-b ity that an energy deficit could explain the observed metabolic
values were attenuated after ET and then recovered after RE dysregulation.
(Figure 3F). In contrast, C-peptide was unchanged throughout
the intervention (Figure 3G), indicating a higher hepatic clear- Excessive training attenuates the glucoregulatory
ance of insulin in the ET phase. response to intense exercise despite increased skeletal
Exercise training affects glucose uptake partly by increasing muscle glycogen stores
the abundance of the glucose transporter 4 protein (GLUT4), High-intensity exercise is enabled through a parallel activation of
which is located in the cytosol during rest and is transported to both oxidative and glycolytic energy pathways. Interestingly,
the membrane upon insulin stimulation (Richter and Hargreaves, lactate has recently been shown to be a primary circulating
2013). We therefore measured the abundance of GLUT4 in mus- fuel that can be used as a carbon source in the tricarboxylic
cle biopsy samples collected before glucose intake and found acid cycle and thereby constitutes a route for the distribution
that, contrary to the impaired glucose uptake in ET, GLUT4 in of mitochondrial fuel across the circulation (Hui et al., 2017).
whole muscle increased and was most abundant after ET (Fig- Therefore, we analyzed blood glucose and lactate during exer-
ure 3H). We also measured the fractional abundance of mem- cise throughout the intervention.
brane-bound GLUT4 but no differences were found between During submaximal work, glucose was not significantly
the intervention phases (Figure 3I). The phosphorylation of Akt affected (Figure 4A), and during HIIT, both blood glucose and
substrate of 160 kDa (AS160) has been implicated in the translo- lactate increased as expected compared with submaximal in-
cation of GLUT4 to the cell membrane upon insulin stimulation tensity. This increase was, however, sharply attenuated as
(Mı̂inea et al., 2005). In agreement with the data on GLUT4 trans- training load increased and was lowest after ET (Figures 4B
location, we did not find any changes in the total amounts of phos- and 4C). The glycolytic energy contribution, therefore, appeared
phorylated levels of AS160 in our intervention (Figures 3J and 3K). inhibited after ET and is probably the main cause of the staled
The observed dysregulation of glucose control during ET cycling performance in ET. After RE, the metabolic switch back
might be a consequence of reduced glucose uptake and/or by to a more glycolytic mode during HIIT was remarkable with the
a reduced capacity to oxidize glucose. The resting metabolic highest measurements of blood glucose and lactate as perfor-
rate (RMR) before glucose load was not different between mance peaked during HIIT.
training situations (Figure S4), and the thermogenic effect of The primary source of lactate originating from glycolysis is
the glucose load increased the metabolic rate by 8% but was skeletal muscle glycogen. To determine whether the blunted
not different in any of the training situations. However, fat and lactate response to HIIT was an effect of the reduced availability
carbohydrate oxidation varied throughout the intervention with of glycogen in the skeletal muscles, we analyzed glycogen con-
the highest measurements made after LT and ET (Figure S5). centrations in muscle tissue from vastus lateralis. Surprisingly,
The change in substrate oxidation can, however, not explain we found that glycogen was only reduced between BL and LT,
the observed shift in glucose tolerance between LT and ET as which probably is explained by the glycogen usage during the
substrate oxidation was unaltered between those time points. first HIIT session and the limited time for repletion. Thereafter,

(J and K) Total AS160 (J) and phosphorylated AS160 (K) were measured by the western blot technique.
(L) Fat oxidation during exercise at submaximal intensity (100 W) was calculated using indirect calorimetry.
(M) Pictures of protein staining for purity of fractions and a reference memcode staining.
For insulin, HOMA-IR and HOMA-b, n = 10; GLUT4 membrane fraction, n = 8. All other measurements, n = 11.

962 Cell Metabolism 33, 957–970, May 4, 2021


ll
Article

Figure 4. Excessive training attenuates the glucoregulatory response to intense exercise despite increased skeletal muscle glycogen stores
(A) Blood glucose was measured after a submaximal work rate.
(B and C) Blood glucose (B) and lactate (C) after the last HIIT interval.
(D) Muscle glycogen content was measured in muscle homogenate from vastus lateralis using a spectrophotometric approach
(E) The activity of hexokinase was measured spectrophotometrically throughout the training period.
(F) Glycogen synthase was measured by the western blot technique.
(G) Fasting plasma free fatty acids were measured spectrophotometrically throughout the training period.
(H) A reference memcode staining.
For all measurements, n = 11.

glycogen increased throughout the different training phases (Fig- FFA were low and remained unaltered throughout the training
ure 4D). Further, we analyzed two key proteins involved in the period, and thus, cannot explain the deterioration in glucose
synthesis of glycogen from glucose. The maximal activity of tolerance or the decreased plasma lactate levels during HIIT in
hexokinase, which catalyzes the conversion of glucose to the ET phase (Figure 4G).
glucose-6-phosphate, was increased after ET (Figure 4E) in Overall, ET appears to inhibit cycling performance through a
agreement with the higher glycogen depots at that time point. blunted glycolytic response to intense exercise, despite normal
Likewise, the amount of glycogen synthase protein increased af- muscle glycogen levels and unaltered metabolic stimulation
ter ET (Figure 4F). Thus, the reduced blood lactate concentration through plasma FFA supply.
measured in ET cannot be explained by reduced glycogen avail-
ability, which previously has been suggested to be linked to Mitochondrial fusion and fission respond to exercise
staled performance during intensified training (Costill et al., training in a dose-dependent manner but are not
1988). This indicates that other metabolic changes likely must specifically modulated after excessive exercise
be important in the regulation of substrate use and performance After moderate exercise training, mitochondrial dynamics have
during intense training. been shown to be activated by the promotion of fusion and
According to the Randle effect (Randle et al., 1963), a stimu- reduction of fission events (Memme et al., 2021). The importance
lated lipid metabolism through elevated levels of blood FFAs of a functional mitochondrial reticulum is gaining increasing
should potentially inhibit glycolytic activity. As elevated circu- attention as an important factor in explaining the health out-
lating FFAs also are associated with insulin resistance we comes of exercise training (Tanaka et al., 2020). For example,
measured blood FFA after overnight fasting in our subjects mitochondrial fragmentation and fission have been implicated
throughout the training period. However, the concentrations of in insulin resistance in skeletal muscle cells (Jheng et al.,

Cell Metabolism 33, 957–970, May 4, 2021 963


ll
Article

Figure 5. Mitochondrial fusion and fission respond to exercise training in a dose-dependent manner but are not specifically modulated dur-
ing excessive exercise
The abundance of mitochondrial outer membrane fuser MFN2 (A), inner membrane fuser OPA1 (B), fission executor DRP1 (C), docking protein MFF (D), and FIS1
(E) all assessed using the western blot technique. Pictures of a reference memcode staining are presented in (F). For all measurements, n = 11.

2012), whereas overexpression of the fusion factors MFN 1 and der normal circumstances. Despite the increase in training load,
MFN 2 has an insulin-sensitizing effect (Lin et al., 2018). we found no significant effect on any of the antioxidative proteins
We found that Mitofusin 2 (MFN2) (Figure 5A), which fuses the SOD2, GPX, CAT, and HO-1-1, which all remained unchanged
mitochondrial outer membrane, as well as inner membrane fuser with training as assessed by western blotting (Figures 6A–6D).
optic atrophy 1 (OPA1) (Figure 5B) both increased robustly dur- To understand how increasing doses of exercise affected the
ing the first 3 weeks of progressive training load and were most overall oxidation status in the skeletal muscles, we measured the
abundant after ET. The activation of mitochondrial fission did not amount of peroxidized lipids and carbonylated proteins in mus-
follow a uniform pattern and the protein content of DRP1 was not cle homogenates taken throughout the study. We expected the
significantly altered throughout the training period (Figure 5C). In level of oxidative stress to increase with increasing exercise
contrast, the abundance of the DRP1 receptor MFF decreased training load. Surprisingly, we found that the total amount of per-
during the intervention (Figure 5D), whereas FIS1 increased oxidized lipids (Figure 6E) and carbonylated proteins (Figure 6F)
and was highest after ET (Figure 5E). were kept unchanged during the intervention.
In human skeletal muscle, the activation of autophagy seems ROS produced during exercise has been shown to originate
to be dependent on exercise intensity, duration, as well as en- from xanthine oxidoreductase, NADPH oxidase, and the mito-
ergy availability (Martin-Rincon et al., 2018). We assessed three chondria themselves (Henrı́quez-Olguin et al., 2019). Knowing
canonical markers of autophagy. The microtubule-associated that the overall oxidation status was maintained despite
protein 1 light chain (LC3BI), which creates docking sites for au- increasing training loads, we wanted to test if mitochondrial
tophagolysosomes and lysosomes, and the lipidated form hydrogen peroxide production was altered in different training
LC3BII, were unchanged throughout the training intervention phases.
(Figures S7A and S7B). Beclin1, an important protein in the auto- We found that the production of H2O2 closely followed
phagosome formation, increased during the intervention and changes in mitochondrial respiration and was reduced in ET
reached a peak value during ET (Figure S7C). compared with the highest values in LT and MT. H2O2 production
Taken together, the presumed scheme of healthy remodeling was lower when both mitochondria respired on complex I sub-
and activation of autophagy appears to be activated in parallel strates alone (Figure 6G) and in a combination of complex I
with increasing exercise volume and cannot explain the dramatic and II substrates (Figure 6H). This decrease in mitochondrial
reduction in IMR during the ET phase. H2O2 production could be a compensatory mechanism to coun-
teract increases in non-mitochondrial ROS production as a pro-
Oxidative stress is maintained during excessive training tective strategy against oxidative stress.
owing to reduced mitochondrial H2O2 emission We have previously reported the redox-sensitive TCA cycle
ROS increase during exercise (Henrı́quez-Olguin et al., 2019) but enzyme, aconitase, to decrease by over 50% in parallel with
are effectively quenched by the antioxidant system that includes mitochondrial respiration after sprint training (Larsen et al.,
mitochondrial superoxide dismutase (SOD2), catalase (CAT), 2016). Here, we found that the mitochondrial aconitase activity
glutathione peroxidase (GPX), and hemeoxygenase (HO-1-1) un- changed during the intervention with a small increase from BL

964 Cell Metabolism 33, 957–970, May 4, 2021


ll
Article

Figure 6. Oxidative stress is maintained during excessive training owing to reduced mitochondrial H2O2 emission with a simultaneous loss
of Nrf2
(A–D) The abundance of proteins in the mitochondrial antioxidative defense system; SOD2 (A), GPX (B), CAT (C), and HO-1-1 (D) were measured using the western
blot technique.
(E) The amount of peroxidized lipids was measured with a TBARS assay.
(F) Carbonylated proteins were measured using a colorimetric approach.
(G and H) Mitochondrial H2O2 production while mitochondria respired on complex I substrates (G) and on both complex I+II substrates (H) were measured using
Amplex UltraRed and horseradish peroxidase in the high-resolution respirometer.
(I) Activity of TCA cycle enzyme aconitase was measured in isolated mitochondria using a colorimetric approach.
(J–L) The abundance of Nrf2 protein (J) and the abundance of Nrf2 retainer protein Keap1 (K) was measured using the western blot technique. The ratio between
Nrf2 and Keap1 is presented in (L).
(M) A reference memcode staining.
A repeated-measure one-way ANOVA was used and a significant main effect (of time) is marked with a # in graphs. Dunnett’s post hoc test was used on normally
distributed data and Dunn’s post hoc test was used on non-normal distributed data. (G) and (H) were assessed using a mixed-model analysis due to missing data.
Post hoc tests were performed between ET and all other situations with the exception of (J)–(L), where an uncorrected Dunn’s test was performed between ET and
MT (annotated with $ in J and K). (I) n = 10; for all other measurements, n = 11.

Cell Metabolism 33, 957–970, May 4, 2021 965


ll
Article
Figure 7. Elite endurance athletes have
disturbed glucose control
Elite endurance athletes and a matched control
group used CGMs in periods up to 2 weeks.
(A–C) The number of minutes per 24-h period that
blood glucose was above 8 mmol$L 1 (A) between
4 and 8 mmol$L 1 (B) and below 4 mmol$L 1 (C).
(D) Mean glucose for athletes and controls over 24
h. Statistical analysis was done using an unpaired
Student’s t test. Athletes, n = 15; controls, n = 12.

after ET (Figure 6L; ANOVA main effect,


p = 0.24; uncorrected Dunn’s test, p =
0.08, MT compared with ET). Nrf2 protein
is continually synthesized and retained in
the cytosol by KEAP1; a decreased Nrf2/
KEAP1 ratio indicates less amounts of
protein that can translocate to the nucleus
where it binds to promoter regions of its
target genes. However, we found no
detectable amounts of Nrf2 in the nucleus
at any time point, possibly reflecting the
timing of the biopsy 14 h after the last ex-
ercise session.

Elite endurance athletes have a


disturbed glucose control
Knowing that ET has detrimental effects
on glucose tolerance and deranges the
glycolytic response to exercise, we
wanted to know if this was a problem in
a real-life setting. To test if the impair-
to MT and was followed by a reduction after ET (Figure 6I). The ments in glucose tolerance found in this study were mirrored
decreased mitochondrial respiration could be interpreted as a as an impaired glucose control in elite athletes, we recruited
protective strategy to maintain redox homeostasis. a second cohort of 15 healthy endurance athletes representing
national teams in endurance sports and measured their glucose
Mitochondrial functional impairment and decreased control every 15th min for up to 2 weeks using continuous
glucose tolerance coincide with loss of Nrf2 glucose monitors (CGMs). Their results were compared with
Having established that mitochondrial respiration and H2O2 age, sex, and a weight-matched control group of 12 subjects
emission decreased after the ET phase, we wanted to find a that trained less than 7 h per week. For subject characteristics,
mediator of this effect. Nrf2 is the master regulator of several see Table S2.
hundred genes involved in the antioxidant defense but knock- We found that in free-living conditions, where athletes and con-
down of Nrf2 has also been shown to reduce mitochondrial trols maintained their normal training and diet, the mean 24-h
respiration and cellular ATP levels (Holmström et al., 2013). Like- glucose was nearly identical in athletes and controls (5.5 ± 0.1
wise, Nrf2 seems to be dysregulated in diabetic patients (Rab- versus 5.5 ± 0.1 mmol$L 1). Surprisingly, the time spent outside
bani et al., 2019), and mice lacking Nrf2 protein have disturbed of the normoglycemic range (4–8 mmol$L 1) was significantly
glucose homeostasis with insulin resistance after a period of ex- higher in athletes compared with controls. Athletes spent 41 min
ercise (Cox et al., 2018). Nrf2 abundance showed a rather high per 24 h in the hyperglycemic range (>8 mmol$L 1) compared
intra-individual variability when the training load was low (in BL with 22 min for the controls (Figure 7A) (t test, p = 0.005),
and RE), and thus, no significant main effect of time was found 1,284 min in the normoglycemic range (4–8 mmol$L 1) compared
(p = 0.16, ANOVA). However, Nrf2 abundance decreased be- with 1,384 min for the controls (Figure 7B) (t test, p = 0.001), and
tween the MT and ET phases, and an uncorrected Dunn’s test 115 min in the hypoglycemic range (<4 mmol$L 1) compared
between these two situations in isolation showed a significant ef- with 34 min for the controls (Figure 7C) (t test, p = 0.04). In the ath-
fect (Figure 6J; p < 0.05). Similarly, the repressor KEAP1 reached letes, hyperglycemia seemed to occur in the early afternoon and
its highest value after ET (Figure 6K; ANOVA main effect, p = hypoglycemia mainly occurred during sleep between 3 and 7
0.12; uncorrected Dunn’s test, p < 0.05, ET compared with a.m. (Figure 7D). Interestingly, normoglycemia seemed to be
MT). Therefore, the Nrf2/KEAP-1 ratio reached its lowest value maintained in most cases even during extensive training sessions.

966 Cell Metabolism 33, 957–970, May 4, 2021


ll
Article

DISCUSSION but is rather an intrinsic mitochondrial defect, as shown in pa-


tients with T2D (Mogensen et al., 2007).
Here, we show that there is an upper limit of the amount of inten- We hypothesized that during excessive exercise the ROS pro-
sive exercise that can be performed without disrupting metabolic duction should increase and overwhelm the endogenous antiox-
homeostasis. Beyond this limit, negative effects on metabolic idative systems. Surprisingly, we found that both the number of
health and adaptation of physical performance, seemingly carbonylated proteins and TBARS were kept unchanged
caused by a mitochondrial partial shutdown of both respiration throughout the study, indicating that the total level of ROS pro-
and H2O2 production, start to manifest. The results of this study duction was constant. In the ET phase, mitochondrial H2O2 pro-
also provide valuable insight into the ongoing controversy about duction was reduced by more than 50% compared with MT. We
whether mitochondrial dysfunction is a cause or occurs second- interpret this finding as a compensatory partial shutdown of
ary to insulin resistance (Goodpaster, 2013; Holloszy, 2013). The mitochondrial metabolism to reduce overall ROS production
decreased IMR and glucose intolerance coincided with an unex- and maintain a balanced redox environment. The reduction in
pected loss of Nrf2 protein. IMR can only partly be explained by the inactivation of the
The exercise protocol used here with progressively redox-sensitive aconitase. When mitochondrial respiration was
increasing training loads is a unique model allowing us to assessed with the complex II substrate succinate together with
examine the relationship between mitochondrial dysfunction, the complex I inhibitor rotenone, aconitase was bypassed but
metabolic health, and glucose intolerance in the same healthy the magnitude of reduction in IMR in the ET phase was only partly
subjects over a short period of time. We found that changes attenuated compared with complex I respiration. Inhibition of
in IMR were closely paralleled by changes in glucose tolerance mitochondrial respiration despite an increase in mitochondrial
(Figure 3C). These results support the hypothesis that mito- proteins has previously been found in a model with transient
chondrial dysfunction indeed causes disturbances in glucose knockdown of SOD2 during embryonic development, where it
metabolism. The classical explanation for such impairment is was interpreted as an adaptive response to oxidative stress
that dysfunctional mitochondria have a lower capacity to (Cox et al., 2018).
oxidize lipid-based substrates that leads to pathological lipid The training load during the ET phase consisted of almost daily
accumulation, which, in turn, is intimately associated with insu- intervals at the all-out effort and must be considered a rather
lin resistance (Morino et al., 2006). In this study, such an expla- extreme type of exercise training that only highly motivated indi-
nation is unlikely, primarily due to the fact that fat oxidation at a viduals can tolerate for more than a few days in a row. Therefore,
submaximal work rate (100 W) increased throughout the we do not see a high risk that people wishing to improve their
training phases, regardless of the drastic impairment in IMR, health through exercise enter the state of mitochondrial impair-
including mitochondrial fat oxidation, in the ET phase. The ment and glucose intolerance due to ET volumes. However, our
rate of fat oxidation during the submaximal work rate was 3.7 finding that the cohort of elite endurance athletes has a disturbed
times higher than at rest, indicating that the mitochondrial ca- glucose control compared with healthy controls is in line with the
pacity for fat oxidation should be far more than the demand hypothesis that ET can cause glucose intolerance. This finding is
during resting conditions. In conjugate, these results suggest highly surprising, knowing that athletes with high oxidative capac-
a dissociation between whole-body fat oxidation and the ity are assumed to have high insulin sensitivity, and during aging,
maximum mitochondrial capacity for fat oxidation and suggest they show less decline in metabolic health than controls (Borgh-
that other factors contribute to the metabolic disturbances outs and Keizer, 2000). To the best of our knowledge, only one
observed in ET. study has assessed continuous glucose profiles in athletes,
The interplay between IMR and insulin sensitivity is complex. wherein ~40% of the subjects were found to spend a substantial
In a study comparing healthy controls with offspring of patients amount of time in the hyperglycemic range (Thomas et al., 2016).
with T2D, exercise training increased mitochondrial ATP produc- The negative metabolic effects of excessive exercise found in this
tion in both groups but only improved insulin sensitivity in the study are likely reduced as soon as the training load decreases.
controls (Irving et al., 2011). In another study, obese insulin- Indeed, we found a substantial improvement in metabolic param-
resistant women had similar IMR but higher mitochondrial eters after only 1 week of reduced training in the RE compared
H2O2 emission than a lean control group. After 12 weeks of with the ET phase.
training, mitochondrial H2O2 emission decreased in the obese From a health perspective, we do not advise against intensive
group in parallel with an increased insulin sensitivity, whereas exercise training as former elite athletes have lower mortality
these parameters were unchanged in the control group (Ko- rates and seem to live longer compared with the general popula-
nopka et al., 2015). tion (Ruiz et al., 2014). Nevertheless, both athletes and those
The reduction in IMR in the ET phase contrasts the response in looking to improve their health through exercise should carefully
markers of mitochondrial density. CS enzyme activity, as well as monitor the response to training, as too much exercise might
mitochondrial protein abundance, including VDAC1 and Mitofi- have negative effects. Using changes in glucose tolerance or
lin, all peaked during ET. This indicates that the ET phase careful tracking of glucose homeostasis using CGM could be a
activated mitochondrial biogenesis pathways and more mito- minimally invasive and a novel approach to optimize the amount
chondrial proteins were synthesized, but they had a lower IMR. of exercise associated with the greatest benefits.
Therefore, the mitochondrial dysfunction observed in this study
does not seem to involve a disturbance in the biogenesis of mito- Limitations of study
chondria or a decreased mitochondrial number, as shown in We used healthy, active subjects and do not know how seden-
some early studies on insulin resistance (Kelley et al., 2002), tary subjects or patients with metabolic disease respond to

Cell Metabolism 33, 957–970, May 4, 2021 967


ll
Article

increasing doses of HIIT exercise. In this cohort, 90 min of HIIT AUTHOR CONTRIBUTIONS
per week was well tolerated, whereas 152 min per week was M.F., L.C.N., S.T., W.A., and F.J.L. performed experiments. M.F. and L.C.N.
associated with maladaptations. The amount of training associ- supervised training sessions. M.F., L.C.N., W.A., and F.J.L. analyzed data.
ated with optimal adaptations is likely lower in subjects not W.A. and B.E. performed muscle biopsies and B.E. had the medical responsi-
accustomed to exercise training, whereas elite athletes should bility. F.J.L. conceived the hypothesis, designed the study, and finalized the
tolerate a higher training load after years of training. Another lim- manuscript. M.F. wrote the first draft and finalized the manuscript. All authors
have read, commented on, and approved the final version of the manuscript.
itation is the timing of the biopsies and venous blood samples,
which were taken ~14 h after the last training session. It is
possible that IMR, mitochondrial H2O2 emission, and blood DECLARATION OF INTERESTS
FFA showed an acute response to each exercise session with F.J.L. is co-founder of Silicon Valley Exercise Analytics, a company using data
more reliance on fat oxidation and more severe impairments of science to improve athletic performance. The company had no role in funding,
IMR immediately after each exercise session and that these pa- data collection and analysis, or preparation of the manuscript. All other authors
rameters had partly recovered by the time point of our biopsy declare no conflict of interest.
collection.
Received: July 12, 2020
Revised: November 18, 2020
STAR+METHODS Accepted: February 22, 2021
Published: March 18, 2021

Detailed methods are provided in the online version of this paper


REFERENCES
and include the following:
Bhatti, J.S., Bhatti, G.K., and Reddy, P.H. (2017). Mitochondrial dysfunction
d KEY RESOURCES TABLE
and oxidative stress in metabolic disorders - a step towards mitochondria
d RESOURCE AVAILABILITY based therapeutic strategies. Biochim. Biophys. Acta Mol. Basis Dis. 1863,
B Lead contact 1066–1077.
B Materials availability
Borg, G.A. (1982). Psychophysical bases of perceived exertion. Med. Sci.
B Data and code availability Sports Exerc. 14, 377–381.
d EXPERIMENTAL MODEL AND SUBJECT DETAILS Borghouts, L.B., and Keizer, H.A. (2000). Exercise and insulin sensitivity: a re-
B Subjects view. Int. J. Sports Med. 21, 1–12.
B Pre-test Boushel, R., Gnaiger, E., Schjerling, P., Skovbro, M., Kraunsøe, R., and Dela,
B Training intervention F. (2007). Patients with type 2 diabetes have normal mitochondrial function in
B Nutritional intervention skeletal muscle. Diabetologia 50, 790–796.
B Continues glucose monitoring Brouwer, E. (1957). On simple formulae for calculating the heat expenditure
d METHOD DETAILS and the quantities of carbohydrate and fat oxidized in metabolism of men
B Muscle biopsies and animals, from gaseous exchange (oxygen intake and carbonic acid
B Oral glucose tolerance test and resting metabolism output) and urine-N. Acta Physiol. Pharmacol. Neerl. 6, 795–802.

B Mitochondrial isolation and respirometry Costill, D.L., Flynn, M.G., Kirwan, J.P., Houmard, J.A., Mitchell, J.B., Thomas,
R., and Park, S.H. (1988). Effects of repeated days of intensified training on
B Immunoblotting
muscle glycogen and swimming performance. Med. Sci. Sports Exerc. 20,
B Free fatty acids
249–254.
B Insulin
Cox, C.S., McKay, S.E., Holmbeck, M.A., Christian, B.E., Scortea, A.C., Tsay,
B C-peptide
A.J., Newman, L.E., and Shadel, G.S. (2018). Mitohormesis in mice via sus-
B Muscle glycogen tained basal activation of mitochondrial and antioxidant signaling. Cell
B Hexokinase activity Metab. 28, 776–786.e5.
B CS activity Eijsvogels, T.M.H., Thompson, P.D., and Franklin, B.A. (2018). The ‘‘extreme
B Carbonyl content exercise hypothesis’’: recent findings and cardiovascular health implications.
B TBARS Curr. Treat. Options Cardiovasc. Med. 20, 84.
B Aconitase activity Ekblom, B. (2017). The muscle biopsy technique. Historical and methodolog-
d QUANTIFICATION AND STATISTICAL ANALYSIS ical considerations. Scand. J. Med. Sci. Sports 27, 458–461.
d ADDITIONAL RESOURCES Evans, W.J., Phinney, S.D., and Young, V.R. (1982). Suction applied to a mus-
cle biopsy maximizes sample size. Med. Sci. Sports Exerc. 14, 101–102.
Franklin, B.A., Thompson, P.D., Al-Zaiti, S.S., Albert, C.M., Hivert, M.F.,
SUPPLEMENTAL INFORMATION
Levine, B.D., Lobelo, F., Madan, K., Sharrief, A.Z., Eijsvogels, T.M.H., et al.
(2020). Exercise-related acute cardiovascular events and potential deleterious
Supplemental information can be found online at https://doi.org/10.1016/j.
adaptations following long-term exercise training: placing the risks into
cmet.2021.02.017.
perspective-an update: a scientific statement from the american heart associ-
ation. Circulation 141, e705–e736.
ACKNOWLEDGMENTS Gallego-Selles, A., Martin-Rincon, M., Martinez-Canton, M., Perez-Valera, M.,
Martı́n-Rodrı́guez, S., Gelabert-Rebato, M., Santana, A., Morales-Alamo, D.,
This project was supported by grants to F.J.L. and W.A. as an Early Career Dorado, C., and Calbet, J.A.L. (2020). Regulation of Nrf2/Keap1 signalling in
Research Fellowship from the Swedish Research Council for Sport Science human skeletal muscle during exercise to exhaustion in normoxia, severe
(CIF), Sweden (P2017-0067, P2018-0083, P2019-0062, P2020-0061, and acute hypoxia and post-exercise ischaemia: influence of metabolite accumu-
D2019-0050). lation and oxygenation. Redox Biol. 36, 101627.

968 Cell Metabolism 33, 957–970, May 4, 2021


ll
Article
Gnaiger, E., and Kuznetsov, A.V. (2002). Mitochondrial respiration at low levels Layec, G., Blain, G.M., Rossman, M.J., Park, S.Y., Hart, C.R., Trinity, J.D.,
of oxygen and cytochrome c. Biochem. Soc. Trans. 30, 252–258. Gifford, J.R., Sidhu, S.K., Weavil, J.C., Hureau, T.J., et al. (2018). Acute
Goodpaster, B.H. (2013). Mitochondrial deficiency is associated with insulin high-intensity exercise impairs skeletal muscle respiratory capacity. Med.
resistance. Diabetes 62, 1032–1035. Sci. Sports Exerc. 50, 2409–2417.
Grossbard, L., and Schimke, R.T. (1966). Multiple hexokinases of rat tissues. Lee, D.C., Sui, X., Artero, E.G., Lee, I.M., Church, T.S., McAuley, P.A.,
Purification and comparison of soluble forms. J. Biol. Chem. 241, 3546–3560. Stanford, F.C., Kohl, H.W., 3rd, and Blair, S.N. (2011). Long-term effects of
Halson, S.L., and Jeukendrup, A.E. (2004). Does overtraining exist? An anal- changes in cardiorespiratory fitness and body mass index on all-cause and
ysis of overreaching and overtraining research. Sports Med. 34, 967–981. cardiovascular disease mortality in men: the aerobics center longitudinal
study. Circulation 124, 2483–2490.
Henriksson, K.G. (1979). Semi-open" muscle biopsy technique. A simple
outpatient procedure. Acta Neurol. Scand. 59, 317–323. Lefort, N., Glancy, B., Bowen, B., Willis, W.T., Bailowitz, Z., De Filippis, E.A.,
Brophy, C., Meyer, C., Højlund, K., Yi, Z., and Mandarino, L.J. (2010).
Henrı́quez-Olguin, C., Knudsen, J.R., Raun, S.H., Li, Z., Dalbram, E., Treebak,
Increased reactive oxygen species production and lower abundance of com-
J.T., Sylow, L., Holmdahl, R., Richter, E.A., Jaimovich, E., and Jensen, T.E.
plex I subunits and carnitine palmitoyltransferase 1B protein despite normal
(2019). Cytosolic ROS production by NADPH oxidase 2 regulates muscle
mitochondrial respiration in insulin-resistant human skeletal muscle.
glucose uptake during exercise. Nat. Commun. 10, 4623.
Diabetes 59, 2444–2452.
Holloszy, J.O. (1967). Biochemical adaptations in muscle. Effects of exercise
on mitochondrial oxygen uptake and respiratory enzyme activity in skeletal Leighton, B., Blomstrand, E., Challiss, R.A., Lozeman, F.J., Parry-Billings, M.,
muscle. J. Biol. Chem. 242, 2278–2282. Dimitriadis, G.D., and Newsholme, E.A. (1989). Acute and chronic effects of
strenuous exercise on glucose metabolism in isolated, incubated soleus mus-
Holloszy, J.O. (2013). Deficiency" of mitochondria in muscle does not cause
cle of exercise-trained rats. Acta Physiol. Scand. 136, 177–184.
insulin resistance. Diabetes 62, 1036–1040.
Lin, H.Y., Weng, S.W., Chang, Y.-H., Su, Y.-J., Chang, C.-M., Tsai, C.-J., Shen,
Holmström, K.M., Baird, L., Zhang, Y., Hargreaves, I., Chalasani, A., Land,
F.C., Chuang, J.-H., Lin, T.-K., Liou, C.-W., et al. (2018). The causal role of
J.M., Stanyer, L., Yamamoto, M., Dinkova-Kostova, A.T., and Abramov, A.Y.
mitochondrial dynamics in regulating insulin resistance in diabetes: link
(2013). Nrf2 impacts cellular bioenergetics by controlling substrate availability
through mitochondrial reactive oxygen species. Oxid. Med. Cell. Longev.
for mitochondrial respiration. Biol. Open 2, 761–770.
2018, 7514383.
Hui, S., Ghergurovich, J.M., Morscher, R.J., Jang, C., Teng, X., Lu, W.,
Lowell, B.B., and Shulman, G.I. (2005). Mitochondrial dysfunction and type 2
Esparza, L.A., Reya, T., Le Zhan, Z., Yanxiang Guo, J., et al. (2017). Glucose
diabetes. Science 307, 384–387.
feeds the TCA cycle via circulating lactate. Nature 551, 115–118.
Irving, B.A., Short, K.R., Nair, K.S., and Stump, C.S. (2011). Nine days of inten- Martin-Rincon, M., Morales-Alamo, D., and Calbet, J.A.L. (2018). Exercise-
sive exercise training improves mitochondrial function but not insulin action in mediated modulation of autophagy in skeletal muscle. Scand. J. Med. Sci.
adult offspring of mothers with type 2 diabetes. J. Clin. Endocrinol. Metab. 96, Sports 28, 772–781.
E1137–E1141. Memme, J.M., Erlich, A.T., Phukan, G., and Hood, D.A. (2021). Exercise and
Issurin, V.B. (2010). New horizons for the methodology and physiology of mitochondrial health. J. Physiol. 599, 803–817.
training periodization. Sports Med. 40, 189–206. Mensink, M., Blaak, E.E., van Baak, M.A., Wagenmakers, A.J., and Saris, W.H.
Jeukendrup, A.E., and Wallis, G.A. (2005). Measurement of substrate oxidation (2001). Plasma free fatty acid uptake and oxidation are already diminished in
during exercise by means of gas exchange measurements. Int. J. Sports Med. subjects at high risk for developing type 2 diabetes. Diabetes 50, 2548–2554.
26 (suppl 1), S28–S37. €nen, J., Lane, W.S.,
Mı̂inea, C.P., Sano, H., Kane, S., Sano, E., Fukuda, M., Pera
Jheng, H.F., Tsai, P.J., Guo, S.M., Kuo, L.H., Chang, C.S., Su, I.J., Chang, and Lienhard, G.E. (2005). AS160, the Akt substrate regulating GLUT4 translo-
C.R., and Tsai, Y.S. (2012). Mitochondrial fission contributes to mitochondrial cation, has a functional Rab GTPase-activating protein domain. Biochem. J.
dysfunction and insulin resistance in skeletal muscle. Mol. Cell. Biol. 32, 391, 87–93.
309–319. Miles, J., Glasscock, R., Aikens, J., Gerich, J., and Haymond, M. (1983). A mi-
Kelley, D.E., and Simoneau, J.A. (1994). Impaired free fatty acid utilization by crofluorometric method for the determination of free fatty acids in plasma.
skeletal muscle in non-insulin-dependent diabetes mellitus. J. Clin. Invest. J. Lipid Res. 24, 96–99.
94, 2349–2356.
Moberg, M., Apró, W., Ekblom, B., van Hall, G., Holmberg, H.C., and
Kelley, D.E., He, J., Menshikova, E.V., and Ritov, V.B. (2002). Dysfunction of Blomstrand, E. (2016). Activation of mTORC1 by leucine is potentiated by
mitochondria in human skeletal muscle in type 2 diabetes. Diabetes 51, branched-chain amino acids and even more so by essential amino acids
2944–2950. following resistance exercise. Am. J. Physiol. Cell Physiol. 310, C874–C884.
Konopka, A.R., Asante, A., Lanza, I.R., Robinson, M.M., Johnson, M.L., Dalla Mogensen, M., Sahlin, K., Fernström, M., Glintborg, D., Vind, B.F., Beck-
Man, C., Cobelli, C., Amols, M.H., Irving, B.A., and Nair, K.S. (2015). Defects in Nielsen, H., and Højlund, K. (2007). Mitochondrial respiration is decreased in
mitochondrial efficiency and H2O2 emissions in obese women are restored to skeletal muscle of patients with type 2 diabetes. Diabetes 56, 1592–1599.
a lean phenotype with aerobic exercise training. Diabetes 64, 2104–2115.
Morino, K., Petersen, K.F., Dufour, S., Befroy, D., Frattini, J., Shatzkes, N.,
Konopka, A.R., Castor, W.M., Wolff, C.A., Musci, R.V., Reid, J.J., Laurin, J.L.,
Neschen, S., White, M.F., Bilz, S., Sono, S., et al. (2005). Reduced mitochon-
Valenti, Z.J., Hamilton, K.L., and Miller, B.F. (2017). Skeletal muscle mitochon-
drial density and increased IRS-1 serine phosphorylation in muscle of insulin-
drial protein synthesis and respiration in response to the energetic stress of an
resistant offspring of type 2 diabetic parents. J. Clin. Invest. 115, 3587–3593.
ultra-endurance race. J. Appl. Physiol. 1985 123, 1516–1524.
Morino, K., Petersen, K.F., and Shulman, G.I. (2006). Molecular mechanisms of
Koves, T.R., Ussher, J.R., Noland, R.C., Slentz, D., Mosedale, M., Ilkayeva, O.,
insulin resistance in humans and their potential links with mitochondrial
Bain, J., Stevens, R., Dyck, J.R., Newgard, C.B., et al. (2008). Mitochondrial
dysfunction. Diabetes 55 (suppl 2), S9–S15.
overload and incomplete fatty acid oxidation contribute to skeletal muscle in-
sulin resistance. Cell Metab. 7, 45–56. Nunnari, J., and Suomalainen, A. (2012). Mitochondria: in sickness and in
health. Cell 148, 1145–1159.
Larsen, F.J., Schiffer, T.A., Sahlin, K., Ekblom, B., Weitzberg, E., and
Lundberg, J.O. (2011). Mitochondrial oxygen affinity predicts basal metabolic Pan, D.A., Lillioja, S., Kriketos, A.D., Milner, M.R., Baur, L.A., Bogardus, C.,
rate in humans. FASEB J. 25, 2843–2852. Jenkins, A.B., and Storlien, L.H. (1997). Skeletal muscle triglyceride levels
Larsen, F.J., Schiffer, T.A., Ørtenblad, N., Zinner, C., Morales-Alamo, D., Willis, are inversely related to insulin action. Diabetes 46, 983–988.
S.J., Calbet, J.A., Holmberg, H.C., and Boushel, R. (2016). High-intensity Pedersen, B.K., and Saltin, B. (2015). Exercise as medicine - evidence for pre-
sprint training inhibits mitochondrial respiration through aconitase inactivation. scribing exercise as therapy in 26 different chronic diseases. Scand. J. Med.
FASEB J 30, 417–427. Sci. Sports 25 (suppl 3), 1–72.

Cell Metabolism 33, 957–970, May 4, 2021 969


ll
Article
Pinti, M.V., Fink, G.K., Hathaway, Q.A., Durr, A.J., Kunovac, A., and Hollander, Tanaka, T., Nishimura, A., Nishiyama, K., Goto, T., Numaga-Tomita, T., and
J.M. (2019). Mitochondrial dysfunction in type 2 diabetes mellitus: an organ- Nishida, M. (2020). Mitochondrial dynamics in exercise physiology. Pflug.
based analysis. Am. J. Physiol. Endocrinol. Metab. 316, E268–E285. Arch. 472, 137–153.
Rabbani, P.S., Soares, M.A., Hameedi, S.G., Kadle, R.L., Mubasher, A., Thomas, F., Pretty, C.G., Desaive, T., and Chase, J.G. (2016). Blood glucose
Kowzun, M., and Ceradini, D.J. (2019). Dysregulation of Nrf2/Keap1 redox levels of subelite athletes during 6 days of free living. J. Diabetes Sci.
pathway in diabetes affects multipotency of stromal cells. Diabetes 68, Technol. 10, 1335–1343.
141–155. Tonkonogi, M., and Sahlin, K. (1997). Rate of oxidative phosphorylation in iso-
Randle, P.J., Garland, P.B., Hales, C.N., and Newsholme, E.A. (1963). The lated mitochondria from human skeletal muscle: effect of training status. Acta
glucose fatty-acid cycle. Its role in insulin sensitivity and the metabolic distur- Physiol. Scand. 161, 345–353.
bances of diabetes mellitus. Lancet 1, 785–789. van der Zwaard, S., de Ruiter, C.J., Noordhof, D.A., Sterrenburg, R., Bloemers,
F.W., de Koning, J.J., Jaspers, R.T., and van der Laarse, W.J. (2016). Maximal
Richter, E.A., and Hargreaves, M. (2013). Exercise, GLUT4, and skeletal mus-
oxygen uptake is proportional to muscle fiber oxidative capacity, from chronic
cle glucose uptake. Physiol. Rev. 93, 993–1017.
heart failure patients to professional cyclists. J. Appl. Physiol. 1985 121,
Ruiz, J.R., Fiuza-Luces, C., Garatachea, N., and Lucia, A. (2014). Reduced 636–645.
mortality in former elite endurance athletes. Int. J. Sports Physiol. Perform. Vigelsø, A., Andersen, N.B., and Dela, F. (2014). The relationship between skel-
9, 1046–1049. etal muscle mitochondrial citrate synthase activity and whole body oxygen up-
Samuelsson, H., Moberg, M., Apró, W., Ekblom, B., and Blomstrand, E. (2016). take adaptations in response to exercise training. Int. J. Physiol. Pathophysiol.
Intake of branched-chain or essential amino acids attenuates the elevation in Pharmacol. 6, 84–101.
muscle levels of PGC-1a4 mRNA caused by resistance exercise. Am. J. Westermann, B. (2010). Mitochondrial fusion and fission in cell life and death.
Physiol. Endocrinol. Metab. 311, E246–E251. Nat. Rev. Mol. Cell Biol. 11, 872–884.
Simoneau, J.A., Veerkamp, J.H., Turcotte, L.P., and Kelley, D.E. (1999). Zechner, C., Lai, L., Zechner, J.F., Geng, T., Yan, Z., Rumsey, J.W., Collia, D.,
Markers of capacity to utilize fatty acids in human skeletal muscle: relation Chen, Z., Wozniak, D.F., Leone, T.C., and Kelly, D.P. (2010). Total skeletal
to insulin resistance and obesity and effects of weight loss. FASEB J 13, muscle PGC-1 deficiency uncouples mitochondrial derangements from fiber
2051–2060. type determination and insulin sensitivity. Cell Metab 12, 633–642.

970 Cell Metabolism 33, 957–970, May 4, 2021


ll
Article

STAR+METHODS

KEY RESOURCES TABLE

REAGENT or RESOURCE SOURCE IDENTIFIER


Antibodies
Anti-Heme Oxygenase 1 antibody (HO-1-1) Abcam Cat#ab13248; RRID: AB_2118663
Anti-Glutathione Peroxidase 1 Abcam Cat#ab22604; RRID: AB_2112120
antibody (GPX)
Anti-Catalase antibody - Peroxisome Abcam Cat#ab16731; RRID: AB_302482
Marker (CAT)
Anti-SOD2/MnSOD antibody (SOD2) Abcam Cat#ab13533; RRID: AB_300434
Anti-Mitofusin 2 antibody [6A8] (MFN2) Abcam Cat#ab56889; RRID: AB_2142629
Anti-OPA1 antibody [EPR11057(B)] (OPA1) Abcam Cat#ab157457; RRID: AB_2864313
Anti-DRP1 antibody [EPR19274] (DRP1) Abcam Cat#ab184247
Anti-TTC11/FIS1 antibody [EPR8412] (FIS1) Abcam Cat#ab156865
Anti-MFF antibody [EPR7360] (MFF) Abcam Cat#ab129075; RRID: AB_11155454
Anti-VDAC1 / Porin antibody Abcam Cat#ab154856; RRID: AB_2687466
[EPR10852(B)] (VDAC1)
Anti-Mitofilin antibody [EPR8749] (Mitofilin) Abcam Cat#ab137057
Anti-Glucose Transporter GLUT4 Abcam Cat#ab654; RRID: AB_305554
antibody (GLUT4)
Anti-Keap1 antibody (Keap1) Abcam Cat#ab139729
Anti-Citrate synthetase antibody Abcam Cat#ab129095; RRID: AB_11143209
[EPR8067] (CS)
LONP1 polyclonal antibody Abnova Cat#PAB28578
Beclin-1 (D40C5) Rabbit mAb (Beclin1) Cell Signaling technology Cat#3495; RRID: AB_1903911
LC3B (D11) XP Rabbit mAb (LC3B) Cell Signaling technology Cat#3868; RRID: AB_2137707
Glut4 (1F8) Mouse mAb (GLUT4) Cell Signaling technology Cat#2213; RRID: AB_823508
NRF2 (D1Z9C) XP Rabbit mAb (Nrf2) Cell Signaling technology Cat#12721; RRID: AB_2715528
p70 S6 Kinase Antibody (p70) Cell Signaling technology Cat#9202; RRID: AB_331676
eEF2 Antibody (eEF2) Cell Signaling technology Cat #2332; RRID: AB_10693546
Glycogen Synthase Antibody (GS) Cell Signaling technology Cat #3893; RRID: AB_2279563
AS160 Antibody (AS160) Cell Signaling technology Cat #2447; RRID: AB_2199376
Phospho-AS160 (Thr642) (D27E6) (pAS160) Cell Signaling technology Cat#8881; RRID: AB_2651042
Anti-mouse IgG, HRP-linked Antibody Cell Signaling technology Cat#7076; RRID: AB_330924
Anti-rabbit IgG, HRP-linked Antibody Cell Signaling technology Cat#7074; RRID: AB_2099233
Chemicals, peptides, and recombinant proteins
SuperSignal West Femto Maximum Thermo Scientific Cat#34096
Sensitivity Substrate
Pierce Bovine Serum Albumin Standard Thermo Scientific Cat#23208
Pre-Diluted Set
Pierce(R) 660nm Protein Assay Reagent Thermo Scientific Cat#22660
Restore PLUS Western Blot Stripping Thermo Scientific Cat#464030
Buffer
Pierce Reversible Protein Stain Kit for PVDF Thermo Scientific Cat#24585
Membranes
Critical commercial assays
Subcellular Protein Fractionation Kit for Thermo Scientific Cat#87790
Tissues
C-peptide ELISA Mercodia Cat#10-1136-01; RRID: AB_2750847
Insulin Human ELISA Kit Invitrogen Cat# KAQ1251
(Continued on next page)

Cell Metabolism 33, 957–970.e1–e6, May 4, 2021 e1


ll
Article

Continued
REAGENT or RESOURCE SOURCE IDENTIFIER
Lipid Peroxidation (MDA) Assay Kit Sigma-Aldrich Cat#MAK085
Aconitase Assay Kit Abcam Cat#ab83459
Protein Carbonyl Content Assay Kit Abcam Cat#126287
Software and algorithms
DatLab 5.2 software Oroboros https://www.bioblast.at/index.php/
MitoPedia:_DatLab
ChemiDoc XRS system with Quantity One Bio-Rad https://www.bio-rad.com/en-se/product/
software (version 4.6.3) quantity-one-1-d-analysis-software?
ID=1de9eb3a-1eb5-4edb-82d2-
68b91bf360fb
ChemiDoc MP with Quantity One software Bio-Rad https://www.bio-rad.com/en-se/product/
(version 6.0.1) quantity-one-1-d-analysis-software?
ID=1de9eb3a-1eb5-4edb-82d2-
68b91bf360fb
Graph Pad prism 8.3.1 for Windows Graph Pad https://www.graphpad.com/support/faq/
prism-831-release-notes/

RESOURCE AVAILABILITY

Lead contact
Further information and requests for resources and reagents as well as datasets and protocols should be directed to and will be ful-
filled by the Lead Contact, Filip J Larsen (filip.larsen@gih.se).

Materials availability
This study did not generate new unique reagents.

Data and code availability


This study did not generate/analyze datasets/code.

EXPERIMENTAL MODEL AND SUBJECT DETAILS

Subjects
We recruited six female and five males to take part in the intervention. All subjects that started the training intervention completed the
full protocol. They were all healthy and recreational active and conducted endurance and strength training on regular basis. Exclusion
criteria were commitment to systematic endurance training of more than 5 hours a week as well as regular high intensity interval
training (HIIT). Detailed subject characteristics are presented in Table S1. Female and male subjects were not divided into, or treated
as sub-groups during the intervention or in the reported results. Also, due to the time-span of the intervention and the unknown inter-
ference of hormonal status during the female subject’s menstrual cycle, they started the intervention at different time-points in their
menstrual cycle. Statistical analyses of our primary outcomes, mitochondrial respiration and glucose tolerance, did not reveal any
sex related differences.
As an observational part of the study, a separate cohort of elite endurance athletes and controls were involved for measurement of
continues blood glucose. The elite endurance athletes were competitive at national team level in endurance sports and the group of
weight, age and height matched controls were recreational active but not competitive athletes. The subjects were included in the two
groups based on their training history and involvement in endurance competition. Subject characteristics are presented in Table S2.
All subjects gave their informed consent prior to inclusion in the study. The study protocol was approved in December 2017 by the
regional ethics committee in Stockholm and conformed to the declaration of Helsinki. The first subject was recruited in February 2018
and the training intervention lasted until December 2019 for the last subject. The interventional study was registered as a clinical trial
in February 2021 and was given the clinical trial identifier NCT04753021.

Pre-test
For inclusion and assessment of baseline physiological characteristics during cycling, a pre-test was performed on a SRM ergometer
(Schoberer Rad Messtechnik, SRM, Ju €lich, Germany). For assessment of the physiological response to submaximal work rate a stan-
dard protocol at the institution was used which comprised of a series of five minutes long intervals separated with one minutes of rest
were capillary blood lactate and glucose were taken and analyzed using a Biosen C-Line Clinic (EKF-diagnostics, Barleben, Ger-
many). The work rate was set individually at 80-100 W at the first stage and thereafter increased with 15-30 W per stage until a sub-

e2 Cell Metabolism 33, 957–970.e1–e6, May 4, 2021


ll
Article

stantial blood lactate accumulation was evident. After a short rest, an incremental maximal exercise test was initiated to determine
VO2max. The test started at the work rate of the previous stage, and increased with 20-30 W $ min-1 until fatigue. VO2max was ex-
pressed as the average of the four consecutive highest 10-second long periods, and Wmax as the power output reached before fa-
tigue. Breath by breath sampling of gas exchange were performed during the entire session using an Oxycon Pro device (Erich Jaeger
GmbH, Hoechberg, Germany). Heart rate was measured continuously (Polar Electro OY, Kempele, Finland) and the subjects rated
their perceived exertion using the BORG scale (Borg, 1982) at every stage and at exhaustion. The equipment was calibrated accord-
ing to manufactures instructions.

Training intervention
All physiological tests and HIIT sessions were carefully supervised by highly experienced personnel with extensive background in
physiological and performance testing. Subjects were blinded to their performance during HIIT sessions throughout the intervention
and were instructed to perform all sessions with the ambition to produce the highest possible mean power output. The intervention
period consisted of 14 HIIT-sessions in total all performed at the laboratory under close surveillance and careful monitoring of power
output and heart rate. In addition during six HIIT-sessions, respiratory gas exchange and capillary lactate were monitored with mea-
surements of performance and physiological response to exercise (referred to as HIIT-sessions in the article). The other 8 sessions
are referred to as HIITtraining sessions. The subjects were blinded to power output, cadence and heart rate during all HIIT and HIIT-
training
sessions. During HIIT sessions, VO2, CO2, blood glucose, lactate and rating of perceived exhaustion (BORG) were measured.
All HIIT sessions started with a submaximal warm up of 10 minutes at 100 W and 70 rpm. Substrate metabolism were calculated from
VCO2 and VO2 measurements using the Brouwer equation (Brouwer, 1957). Thereafter five 4 minutes long intervals followed, at
approximately 95% of VO2max intersected with 3 minutes of non-pedaling rest. The ergometer was set in constant brake mode
at an individually determined braking force. The subjects could then alter power output by variation in cadence or by choice of
gear. The subjects were instructed to perform all HIIT sessions to achieve the highest possible mean power output over all intervals.
In order to help the subjects with a recommended (standardized) pacing strategy they were paced at the first interval at a power
output corresponding to their previously highest measured mean power during a HIIT session. The HIITtraining-sessions comprised
of the standardized warm up of 10 min at 100 W and five 8 minutes long intervals at approximately 90% of VO2max intersected
with 3 minutes of non-pedaling rest during the first three weeks of the study. During the fourth week when training load was reduced,
the HIITtraining sessions were shortened to 3x8 minutes and 3x4 minutes (repetitions and minutes). The subjects often trained in pairs
and additional ergometers were used (Monark LT2, Monark Exercise AB, Vansbro, Sweden) besides the SRM during the HIITtraining
sessions. Outside of the training in the laboratory, subjects were only allowed to perform low intensity exercise and strength training
for the upper body. All individually performed training sessions were scheduled to not interfere with performance and physiological
measurements during HIIT sessions.

Nutritional intervention
Energy intake was standardized for each subject during the day of test sessions. Prior to the first HIIT-session, no exercise was
performed but was standardized in terms of rest (48 hour without training) and diet (24 hour). Subjects had their own standardized
breakfast, lunch and additional snacks, which were consumed three hours before reporting to the laboratory in late afternoon for HIIT-
sessions. The meals were individually chosen by the subjects and reflected their normal nutritional intake, and comprised of mix of
carbohydrates, protein and fat and was repeated prior to all biopsies and OGTT. After each HIIT and HIITtraining session a recovery
drink containing 1 g $ kg-1 bw of carbohydrates and 0.25 g $ kg-1 bw of protein was ingested by the subjects. After HIIT sessions,
subjects were supplied with an evening meal consisting of 74 g carbohydrates, 38 g protein and 26 g fat that was consumed two
hours post exercise. They thereafter remained fasted and reported to the laboratory in early morning for an oral glucose tolerance
test (OGTT) and donation of a muscle biopsy. Therefore, each subject had their own individual schedule for strict timing of training
sessions, biopsies, OGTT and diet in order to standardize the last 24 hour prior to each OGTT.

Continues glucose monitoring


In the separate cohort of elite endurance athletes, we continuously assessed blood glucose every 15th minute for up to two weeks
using a skin-mounted sensor (FreeStyle Libre, Abbot) that measure glucose in the interstitial fluid. The sensor was fitted laterally on
the subject’s deltoideus and automatically stored measurements for a maximum of eight hours before it had to be emptied using a
portable reader. If the sensor failed to measure, or if the subject did not export data from the sensor to the reader within eight hour
from the last export, the sensor overwrites previously stored data and leave a gap in the exported text fil. All text-files were manually
checked and abnormal values were excluded. The first 12 hours of measurements were also excluded from all subjects due to dif-
ferences in time before readings appeared stable. In total, 26 687 separate measurements are used for analysis and the mean effec-
tive wear-time was ten days and seven hours for each subject.

METHOD DETAILS

Muscle biopsies
All biopsies were collected in fasted state in early morning following 14 hour of rest after the last HIIT session. Biopsies were taken
from alternating leg in randomized order between subjects from vastus lateralis. First, local anesthesia (2 % Carbocain, AstraZeneca,

Cell Metabolism 33, 957–970.e1–e6, May 4, 2021 e3


ll
Article
€lje, Sweden) was injected at the biopsy site. A small incision was made and approximately 150 mg of wet tissue was removed
Söderta
with a Weil-Blakesley chonchotome (Henriksson, 1979) or a 5 mm Bergström needle with manually applied suction (Evans et al.,
1982). A detailed description of the biopsy technique has previously been described (Ekblom, 2017). The muscle samples were
blotted and dissected clean from visual fat, connective tissue and blood and divided in to three portions; 50 mg to ice cold ISO-
medium for respirometrics and two portions of 50-100 mg to liquid nitrogen and thereafter stored in -80 C for later analysis.

Oral glucose tolerance test and resting metabolism


Each subject completed four OGTTs throughout the study. After arrival to the laboratory resting metabolic rate (RMR) was measured
in supine position by gas exchange measurement using mixing chamber method (Oxycon Pro) attached to a ventilated hood. There-
after a muscle biopsy was taken and a catheter was inserted into the antecubital vein of the forearm and 15 ml blood was drawn. After
10 minutes of passive rest 75 mg glucose dissolved in 300 ml water was ingested and at 0, 15, 30, 45, 60 75, 90, 105- and 120-minutes
after glucose intake a venous blood sample were taken and instantly analyzed for glucose with enzymatic method in the Biosen C-
Line Clinic. At time points 30, 60, 90- and 120-minutes a 15 ml sample of venous blood was taken and kept on ice. The sample were
centrifuged at 2800 rcf and the plasma was pipetted into eppendorf tubes and stored at -80 C until later analysis. Additional RMR
measurements were performed at 40 minutes after glucose ingestion to capture substrate oxidation at the time point when glucose
peaked as well as at the end of OGTT at 115 minutes. A schematic of time-points during OGTT is presented in Figure S1.

Mitochondrial isolation and respirometry


Mitochondria were isolated from 50 mg of muscle in isolation medium (Sucrose 100 mM, KCl 100 mM, Tris-HCl 50 mM, KH2PO4
1 mM, EGTA 100 mM, BSA 0.1%; pH 7.4 as described by Gnaiger and Kuznetsov, 2002). Isolation of mitochondria were performed as
previously described (Tonkonogi and Sahlin, 1997) with modifications (Larsen et al., 2011). Briefly, muscle was kept on ice and ho-
mogenized first by scissors and thereafter after addition of 0.2 mg ml-1 bacterial protease, further homogenization was performed in a
water-cooled glass homogenizer. The homogenate was centrifuged at 700 rcf at 4 C for 10 min in a 15 ml tube and the supernatant
was transferred to new 1.5 ml tubes and centrifuged at 10000 g at 4 C. The pellets were resolved and transferred to a single 1.5 ml
tube and centrifuged at 7000 rcf at 4 C for 5 min giving a resultant pellet that was liquated in 0.6 ml per initial mg wet weight of muscle
in preservation medium (EGTA 0.5 mM, MgCl2 6H2O 3 mM, K-lactobionate 60 mM, Taurine 20 mM, KH2PO4 10 mM, HEPES 20 mM,
Sucrose 110 mM, BSA 1 g L-1 Histidine 20 mM, Vitamin E succinate 20 mM, Glutathione 3 mM, Leupeptine 1 mM, Glutamate 2 mM,
Malate 2 mM, Mg-ATP 2 mM) (Gnaiger and Kuznetsov, 2002).
Mitochondrial respiration and H2O2 emission were measured using a two-channel high-resolution respirometer with an attached
fluorescent probe (Oxygraph-2k, Oroboros Instruments Corporation, Innsbruck, Austria). 10 ul of isolated mitocondria were added to
two 2 ml wells containing respiration medium MIR05 (EGTA 0.5 mM, MgCl2.6H2O 3 mM, K-lactobionate 60 mM, Taurine 20 mM,
KH2P04 10 mM, HEPES 20 mM, Sucrose 110 mM, BSA 1 g L-1). All experiments were performed at 37 C and O2 calibration and
H2O2 standard calibration was performed according to the manufactures’ instructions. The protocol for assessing respiration in
the different states were: Channel A: 1 U/mL Horseradish peroxidase + 10 mM amplex ultrared (H2O2 emission), octanoyl carnitine
0.2 mM + malate 0.5 mM (leak fat), 2.5 mM ADP (fat respiration), rotenone 0.5 mM + succinate 10 mM (Complex II), N2-induced
anoxia/reoxygenation, titration of FCCP 0.05 mM steps (uncoupled respiration). Channel B: 1 U/mL Horseradish peroxidase +
10 mM amplex ultrared (H2O2 emission), pyruvate 5 mM + glutamate 10 mM + malate0.5 mM (Leak PGM), ADP 0.75 mM (complex
I), succinate 10 mM (Complex I+II), N2-induced anoxia/reoxygenation and cytochrome C 10 mM (membrane integrity). Mitochondrial
respiration in all states are related to protein content analyzed in the mitochondrial suspension using spectrophotometric analysis
and Pierce 660 nm protein assay (Thermo Fisher Scientific). All measurements were performed and analyzed in DatLab 5.2 software
(Oroboros, Paar, Graz, Austria).

Immunoblotting
Homogenization was performed in freeze dried samples of 2 mg muscle. The protocol for homogenization is extensively described
earlier (Samuelsson et al., 2016). Briefly, 100 ml/mg dry WT of homogenization buffer consisting of 2 mM HEPES (pH 7.4), 1 mM EDTA,
5 mM EGTA, 10 mM MgCl2, 50 mM b-glycerophosphate, 1% Triton X-100, 1 mM Na3VO4, 2 mM dithiothreitol, 1% phosphatase in-
hibitor cocktail (Sigma P-2850) and 1% (vol/vol) Halt Protease Inhibitor Cocktail (Thermo Scientific, Rockford, IL) was added to each
sample and processed in a bullet blender until homogenized. After rotation for 30 min at 4 C and centrifugation at 10,000 g for 10 min
at 4 C the supernatant was analyzed for protein content and diluted with homogenization buffer and Laemmli buffer (Bio-Rad, Rich-
mond, CA) obtaining a protein concentration of 1 mg/ml. All samples were denatured at 95 C for five minutes and stored at -80 C.
For analyses of cellular fractions a commercial kit was used (Subcellular protein fractionation kit for tissues, Thermo Scientific
87790). 40 mg of wet weight muscle was used for homogenization and treated according to the protocol generating a cytoplasmic
fraction, a membrane fraction and a nuclear fraction. Each fraction was analyzed for protein content and diluted with homogenization
buffer and Laemmli buffer for immunoblotting obtaining a protein concentration of 0.3, 0.06 and 0.25 mg/ml, respectively.
Immunoblotting was performed by the method described extensively in Moberg et al. (2016). Briefly, samples of muscle homog-
enates (20, 16 or 14 mg of protein) were loaded onto 16 or 26-well Criterion TGX gradient gels (4–20% acrylamide; Bio-Rad). Due to
dilution of samples 4.2 mg were loaded of the cytosolic fraction, 0.84 mg of the membrane fraction and 3.5 mg of the nuclear fraction
onto 26-well gels. Electrophoresis (300 V for 32 minutes kept on ice) were performed in transfer buffer containing 25 mM Tris base,
192 mM glycine, and 10% methanol and the proteins were then transferred to a polyvinylidine fluoride membranes (Bio-Rad) at a

e4 Cell Metabolism 33, 957–970.e1–e6, May 4, 2021


ll
Article

constant current of 300 mA for 3 h kept on ice. The membranes were stained with MemCode Reversible Protein Stain Kit (Thermo
Scientific) to be used as a loading control. The membranes were then destaind and blocked with Tris-buffered saline (TBS; 20 mM
Tris base, 137 mM NaCl, pH 7.6) containing 5 % nonfat dry milk for 1 h at room temperature. Thereafter, incubation overnight followed
with antibodies diluted with TBS buffer with 2.5 % non-fat dry milk and 0.1% Tween. The antibodies used were from Abcam; HO-1-1
(ab13248), GPX (ab22604), CAT (ab16731), SOD2 (ab13533), MFN2 (ab56889), OPA1 (ab157457), DRP1 (ab184247), FIS1
(ab156865), MFF (ab129075), VDAC1/Porin (ab154856), Mitofilin (ab137057) GLUT4 (ab654), KEAP1 (ab139729) and CS
(ab129095), from Abnova; LONP1 (PAB28578) and from Cell Signaling Technology; BECLIN1 (3495), LC3B (3868s), GLUT4
(2213), Nrf2 (12721), p70 (9202), eEF2 (2332) GS (3893), AS160 (2447), pAS160 (8881). After incubation, the membranes were washed
and incubated with secondary antibodies conjugated with horseradish peroxidase (Cell Signaling Technology; anti-mouse 7076s,
anti-rabbit 7074s) for 1 h at room temperature. The membranes were washed again and Super Signal West Femto Chemiluminescent
Substrate (Thermo Scientific) was added and target proteins visualized and quantified in Molecular Imager ChemiDoc XRS system
with Quantity One software (version 4.6.3; Bio-Rad) and in ChemiDoc MP with Quantity One software (version 6.0.1; Bio-Rad). When
applicable, membranes were stripped with Restore Western Blot Stripping Buffer (Thermo Scientific) before block in TBS+milk and
incubation with new primary antibodies. All target proteins were loaded so that the proteins of each subject were treated and exposed
to the same conditions as well as visualization and quantification. All proteins from muscle samples are related to eEF2-protein
located at 90-100 kD on the memcode with exception to Nrf2, GLUT4 and Keap1, GS, AS160, eEF2 and p70 which are related to
whole lanes. The cytosolic, membrane and nucleic fractions are all related to whole lanes.

Free fatty acids


Plasma FFA concentration were analyzed enzymatically in blood plasma by letting FFA react with Acyl-CoA synthase and later form
the product NAD+. The formation of NAD+ was measured at 340 nm during 120 minutes in a spectrophotometer (TECAN Infinite 200
Pro) and the amount of FFA’s were calculated. Method description is found in Miles et al. (1983).

Insulin
Plasma levels of insulin were measured using an ELISA kit (Invitrogen, KAQ1251) according to the manufacturer’s instructions.

C-peptide
Plasma levels of C-peptide were measured using an ELISA kit (Mercodia, Uppsala, Sweden, Cat # 10-1136-01) according to the man-
ufacturer’s instructions.

Muscle glycogen
Muscle glycogen was analyzed in 2 mg of freeze dried and dissected muscle using a homogenization protocol described earlier
(Leighton et al., 1989) with small modifications. Briefly, the samples were heated in 1 M KOH (70 C for 20 min) and thereafter incu-
bated in NaAc+HAc and NaAc+amyloglucosidase (40 C for two hour). Reagents and samples were added to cuvettes and incubated
for 20 minutes before reading of absorbance at 340 nM in a spectrophotometer (Beckman Coulter DU800).

Hexokinase activity
The activity of hexokinase was analyzed spectrophotometrically in homogenate from freeze dried muscle. The method has previously
been described (Grossbard and Schimke, 1966). Absorbance was measured over 20 minutes at 340 nm using a spectrophotometer
(TECAN Infinite 200 Pro).

CS activity
Citrate synthase activity were measured in 1-2 mg of freeze dried muscle. Extraction buffer (100ml/mg) containing 50 mM Tris, 1 mM
EDTA, and 5 mM MgCl2, pH 8.2 were added to the samples which were homogenized using a bullet blender. The supernatant was
transferred to new tubes and centrifuged at 1400 rcf rpm for 30 seconds and the pellet was discarded. Cuvettes was loaded with
reaction buffer containing 250 ml 100 mM Tris-HCl, 0.4 mM DTNB, 25 ml 0.1 mM acetyl-CoA 25 ml 1% triton-X, 170 ml H2O and
5 ml sample. Reaction was started with addition of 25 ml 7 Mm oxaloacetate. Absorbance were measured using spectrophotometric
method (Beckman Coulter DU800) at 412 nm at 25  C. CS-activity was also measured in the suspension of isolated mitochondria with
the same reagents.

Carbonyl content
Carbonyl content were analyzed in samples from 1 mg freeze dried muscle dissolved in 1M KOH and heated for 15 minutes at 70  C
using Abcam Protein Carbonyl Content Assay Kit (ab126287). Briefly, DNPH and TCA were added to samples, centrifuged for 5 mi-
nutes at 16 000 rcf and the pellet were dissolved in aceton, put in a bullet blender for mixing and then centrifuged for 2 minutes at 16
000 rcf. The resulting pellet were dissolved in guarnidine solution and added to a 96 wells microplate included in the kit for reading of
absorbance at 360 nm in spectrophotometer (TECAN Infinite 200 Pro). Samples were then moved to new wells and protein content
was determined using Thermo Scientific Pierce Protein Assay Kit (22660) and measured at 660nm.

Cell Metabolism 33, 957–970.e1–e6, May 4, 2021 e5


ll
Article

TBARS
Malondialdehyde (MDA), were analyzed using Sigma Aldrich Lipid Peroxidation Assay Kit (MAK085). The assay is based on the re-
action between MDA and thiobarbituric acid (TBA) forming a colorimetric product proportional to the MDA content. Homogenate (see
CS-activity for homogenate content) containing 1 mg freeze dried muscle was added to MDA standard solution and BHT. TBA-so-
lution were added to samples and incubated at 95  C for 60 minutes before absorbance was measured at 532 nm in a Beckman
Coulter DU800 Spectrophotometer.

Aconitase activity
Aconitase activity were measured in the mitochondrial fraction using assay kit (ab83459, Abcam). Absorbance was measured at
450 nm in a spectrophotometer (TECAN Infinite 200 Pro).

QUANTIFICATION AND STATISTICAL ANALYSIS

Repeated measures one-way ANOVA was used to analyze which of the measured parameters were affected by the progressive
training (main time effect). Normality was assessed using a Shapiro-Wilk test. If a significant main effect was found, Dunnetts
post-hoc test was used that controls for multiple comparisons. In case of non-normally distributed data Friedmans’ test was
used instead of a one-way ANOVA and Dunn’s post-hoc test was used to compare different time points. Post-hoc tests were per-
formed between the ET-phase and all other conditions with exception of Nrf2 and Keap1 (Figure 5) where an uncorrected Dunn’s test
were performed between ET and MT only. Correlations were calculated using Pearson’s correlation coefficient. When only two
groups were compared (elite vs control, Figure 7) unpaired Student’s t tests were used. In case of missing data (Figures 6G and
6H only) a mixed model analysis was utilized. All statistical analyses were performed in Graph Pad prism 8.3.1 for Windows (Graph-
Pad Software, San Diego, California USA) and a p value of <0.05 was considered significant (marked with # for main effect ANOVA,
with * for a significant difference between ET and other situations). Results are expressed as mean ± standard error of the mean
(SEM). Graphical abstract was created using Biorender.

ADDITIONAL RESOURCES

This trial was registered at clinicaltrials.gov with a clinical trial registry number: NCT047653021. The study was performed at Swedish
School of Sport and Health Sciences between February 2018 – February 2021

e6 Cell Metabolism 33, 957–970.e1–e6, May 4, 2021


Cell Metabolism, Volume 33

Supplemental information

Excessive exercise training causes mitochondrial


functional impairment and decreases
glucose tolerance in healthy volunteers
Mikael Flockhart, Lina C. Nilsson, Senna Tais, Björn Ekblom, William Apró, and Filip J.
Larsen
SUPPLEMENTAL TABELS AND FIGURES

Age Weight Height Body fat VO2max Wmax


n= (year) (kg) (cm) (%) (ml · kg-1 · min-1) (W)
Female 6 25 ± 3 61 ± 3 165 ± 1 16.8 ± 0.8 46.7 ± 1.8 233 ± 8
Male 5 31 ± 3 79 ± 3 179 ± 2 12.5 ± 3.0 50.5 ± 1.4 303 ± 18
All 11 27 ± 2 69 ± 3 171 ± 3 14.9 ± 1.5 48.4 ± 1.3 265 ± 14

Supplemental table 1. Subject characteristics in the training intervention. Related to figure 1.


VO2max and Wmax values were measured during cycling at the graded exercise test to fatigue
performed before the intervention started. Values are presented as mean ± S.E.M.

Age Weight Height


n= (year) (kg) (cm)
Athletes Female 6 26 ± 2 57 ± 3 169 ± 2
Control Female 7 27 ± 2 59 ± 2 167 ± 2
Athletes Male 9 29 ± 1 72 ± 2 182 ± 3
Control Male 5 26 ± 2 71 ± 3 176 ± 1

Supplemental table 2. Subject characteristics of endurance athletes and controls from


continuous glucose measurements (CGM). Related to figure 7.

Figure S1. Schematic of OGTT showing timing of biopsies, blood sampling and RMR
measurements. Related to figure 1.
Figure S2. Intrinsic mitochondrial respiration expressed per CS-activity in the same
mitochondrial fraction. Related to figure 2. All values as means ± S.E.M, n = 11.

Figure S3. The abundance of proteins in the electron transport chain for complex I (A), IV
(B), II (C) III (D) and V (E). Pictures of representative protein staining and a reference
memcode staining is presented in (F). A repeated measures one-way ANOVA was used and a
significant main effect (of time) is marked with a # in graphs. Dunnetts’ Post-hoc test was
used on normal distributed data and Dunn’s post-hoc test was used on non-normal distributed
data. Post-hoc tests were performed between ET and all other situations. Related to figure 2.
All values are mean ± S.E.M. For all measurements; n = 11.
Figure S4. Resting Metabolic Rate was measured using indirect calorimetry and a ventilated
hood while the subjects were resting in quiet and thermoneutral conditions. Related to figure
3. All values as means ± S.E.M, n = 11.

Figure S5. The relative ratio of substrate oxidation was calculated from the ratio of exhaled
carbon dioxide and consumed oxygen. Related to figure 3. All values as means ± S.E.M, n =
11.

Figure S6. Body weight measured before every HIIT session remained unchanged during the
intervention. Related to figure 3. All values as means ± S.E.M, n = 11.
Figure S7. The abundance of autophagy proteins LC3BI (A), LC3BII (B) and Beclin1
together with a reference memcode staining (C) were all assessed using western blot
technique. A repeated measures one-way ANOVA was used and a significant main effect (of
time) is marked with a # in graphs. Dunnetts’ Post-hoc test was used on normal distributed
data and Dunn’s post-hoc test was used on non-normal distributed data. Post-hoc tests were
performed between ET and all other situation. Related to figure 5. All values are mean ±
S.E.M. For all measurements; n = 11.

You might also like