ArticlePDF AvailableLiterature Review

The molecular athlete: Exercise physiology from mechanisms to medals

Authors:

Abstract and Figures

Human skeletal muscle demonstrates remarkable plasticity, adapting to numerous external stimuli including the habitual level of contractile loading. Accordingly, muscle function and exercise capacity encompass a broad spectrum, from inactive individuals with low levels of endurance and strength, to elite athletes who produce prodigious performances underpinned by pleiotropic training-induced muscular adaptations. Our current understanding of the signal integration, interpretation and output coordination of the cellular and molecular mechanisms that govern muscle plasticity across this continuum is incomplete. As such, training methods and their application to elite athletes largely rely on a "trial and error" approach with the experience and practices of successful coaches and athletes often providing the bases for "post hoc" scientific enquiry and research. This review provides a synopsis of the morphological and functional changes along with the molecular mechanisms underlying exercise adaptation to endurance- and resistance-based training. These traits are placed in the context of innate genetic and inter-individual differences in exercise capacity and performance, with special considerations given to the ageing athletes. Collectively, we provide a comprehensive overview of skeletal muscle plasticity in response to different modes of exercise, and how such adaptations translate from "molecules to medals".
From evolution to modern-day athleticism. Evolutionary selection of 5 main traits has facilitated the prolonged upright, bounding, bipedal locomotion in humans. Energetic barriers are lowered by long, spring-like tendons (in particular the Achilles tendon), the longitudinal plantar foot arch, ankle ligaments, long legs, in particular femur length, and short toes (to increase stride length and reduce vertical trajectories for better locomotor economy), thinner heart ventricles and larger cavity, increased hind limb muscle mass, and other adaptations. Skeletal strength is conferred, e.g., by larger joint areas in lower but not upper limbs to dissipate impact forces. Stabilization for bipedal movement is mediated by large erector spinae and gluteus muscles opposed to reduced forearm mass and an elongated, narrow waist and broad shoulders to facilitate counterrotation of thorax and arms, while decreased facial length helps head stabilization or an integrated system of bio-tensegrity for embedded perturbation repelling. Eccrine sweat glands, with a particular high density in the head for brain cooling, reduced body hair, dense skin vascularization, mouth breathing, and a large nasal epithelial area all contribute to thermoregulation. Finally, coevolution of locomotion and the brain resulted in expanded cerebello-cerebralcortical circuitry (for anticipation, pre-preparation, sensory integration, pre-planned multilevel compensation to deal with perturbations and destabilizations), cognitive capabilities (to recognize landmarks, long-range orientation, recognizing prey and predators, tracking and speculative tracking/anticipation), perception, fine motor control, and balance. See Refs. 1-8 for more information. Modern-day athletic peak performances most likely exceed these general evolutionary traits because of efficient training strategies and paradigms, nutrition and supplements, technological innovations, i.e., pertaining to equipment and facilities, and genetic and epigenetic predispositions. Figure created with BioRender.com, with permission.
… 
Content may be subject to copyright.
REVIEW ARTICLE
THE MOLECULAR ATHLETE: EXERCISE
PHYSIOLOGY FROM MECHANISMS TO MEDALS
AUTHORS
Regula Furrer, John A. Hawley,
Christoph Handschin
CORRESPONDENCE
christoph.handschin@unibas.ch;
regula.furrer@unibas.ch;
john.hawley@acu.edu.au
KEY WORDS
athlete; endurance training; exercise; resistance
training; skeletal muscle
CLINICAL HIGHLIGHTS
1) During human evolution, Homo sapiens emerged as mobile hunters and gatherers, dependent on the natural avail-
ability of food. However, todays sedentary lifestyle and overabundant food availability place a major burden on our
metabolic health and are strong drivers underpinning the dramatic rise in noncommunicable diseases.
2) A sedentary lifestyle, characterized by low maximal oxygen uptake (V
.O
2max
), unfavorable body composition, and low
muscle strength, is an independent risk factor for many chronic diseases and a strong predictor of morbidity and mor-
tality.
3) Despite marked interindividual differences in the response to standardized exercise training, regular physical activity
lowers the risk of and confers therapeutic benets for many noncommunicable diseases.
4) Endurance- and resistance-based exercise training protocols confer distinct clinical and health-related benets and
can prevent or reverse many lifestyle-induced metabolic diseases.
5) Clinical exercise tests based on established, validated physiological outcomes are essential for the diagnosis and
subsequent monitoring of clinical conditions.
6) Investigations of elite human performance provide valuable insights into the molecular, cellular, tissue, and whole
body adaptations to extreme metabolic loading. Identication of the mechanisms and pathways that limit exercise
capacity may ultimately aid in the identication of novel therapeutic targets to be prescribed to patient populations.
FURRER ET AL., 2023, Physiol Rev 103: 16931787
January 5, 2023; Copyright © 2023 The Authors. Licensed under Creative Commons Attribution CC-BY 4.0.
Published by the American Physiological Society.
https://doi.org/10.1152/physrev.00017.2022
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
THE MOLECULAR ATHLETE: EXERCISE
PHYSIOLOGY FROM MECHANISMS TO MEDALS
Regula Furrer,
1
John A. Hawley,
2
and Christoph Handschin
1
1
Biozentrum, University of Basel, Basel, Switzerland and
2
Exercise and Nutrition Research Program, Mary MacKillop Institute for
Health Research, Australian Catholic University, Melbourne, Victoria, Australia
Abstract
Human skeletal muscle demonstrates remarkable plasticity, adapting to numerous external stimuli including the
habitual level of contractile loading. Accordingly, muscle function and exercise capacity encompass a broad spec-
trum, from inactive individuals with low levels of endurance and strength to elite athletes who produce prodigious
performances underpinned by pleiotropic training-induced muscular adaptations. Our current understanding of the
signal integration, interpretation, and output coordination of the cellular and molecular mechanisms that govern
muscle plasticity across this continuum is incomplete. As such, training methods and their application to elite ath-
letes largely rely on a trial-and-errorapproach, with the experience and practices of successful coaches and ath-
letes often providing the bases for post hocscientic enquiry and research. This review provides a synopsis of
the morphological and functional changes along with the molecular mechanisms underlying exercise adaptation to
endurance- and resistance-based training. These traits are placed in the context of innate genetic and interindivid-
ual differences in exercise capacity and performance, with special consideration given to aging athletes.
Collectively, we provide a comprehensive overview of skeletal muscle plasticity in response to different modes of
exercise and how such adaptations translate from molecules to medals.
athlete; endurance training; exercise; resistance training; skeletal muscle
1. INTRODUCTION AND BACKGROUND 1693
2. OPTIMIZING TRAINING ADAPTATIONS TO... 1699
3. PHYSIOLOGICAL AND CELLULAR... 1709
4. ACUTE MOLECULAR MECHANISMS... 1722
5. CAN WE ALL BECOME GOLD MEDALISTS? 1746
6. SUMMARY, CONCLUSIONS, AND... 1754
1. INTRODUCTION AND BACKGROUND
1.1. Historical Context: The Evolution of Human
Movement
The evolution of humankind is inextricably linked to the
attainment of an upright, bipedal gait, which conferred
an advantage for locomotion, foraging, and recognition
of prey and predators (FIGURE 1). Indeed, a superior en-
durance capacity, coupled with an outstanding ability to
thermoregulate, was essential for human survival (1).
Evolutionary theory describes the mechanism of natural
selection as survival of the ttest,the underlying sup-
position being that the t,as opposed to the unt,
had a greater likelihood of survival (9). In this regard,
human skeletal muscles, limbs, and the supporting venti-
latory, cardiovascular, and metabolic systems were well
suited for upright locomotion, with economy of move-
ment for bipedal walking and running far exceeding that
of other primates (25). Modications in bone and carti-
lage structure, larger limbs and joints, and spring-like
plantar arches (2,3), combined with a robust system of
perception, ne motor control, and balance, were linked
to a larger brain size and associated cognitive sophisti-
cation (68). The evolution of the larger brain in humans
was likely facilitated by the running behavior of our
ancestors that enabled the procurement of high-pro-
tein sources of food essential for brain development
(10). Bipedal, long-distance running not only necessi-
tates complex computation to control gait, balance,
and stride but also requires large-scale cognitive
processes to recall landmarks associated with abun-
dant sources of food, recognize prey and predators,
and enable long-range orientation (10). Such adapta-
tions were supported by adequate energy availability
andoxygensupply,coupledwithahighdegreeof
metabolic regulation and exibility (68). The superior
human prociencies as hunters, gatherers, and ulti-
mately farmers provided dietary subsistence that
enabled the evolution of our energy-costly brain. The
coevolution of skeletal muscle and associated organ
systems was characterized by progressive and itera-
tive mutual interactions (10). The behavioral lifestyle
and energy availability were determined by the
0031-9333/23 Copyright © 2023 The Authors. Licensed under Creative Commons Attribution CC-BY 4.0.
Published by the American Physiological Society.
1693
Physiol Rev 103: 16931787, 2023
First published January 5, 2023; doi:10.1152/physrev.00017.2022
REVIEW ARTICLE
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
periodic cycles of feasts and famines, with certain
genes evolving to regulate efcient storage and utili-
zation of endogenous fuel stores, the so-called thrifty
genes(11,12).
In contrast to the strong evolutionary pressure to opti-
mize endurance capacity (1), the control of skeletal mus-
cle mass and strength evolved in a more restrained
manner.Althoughadequatemusclestrengthwasclosely
aligned to the prevailing environmental demands of the
day and was indispensable for survival, genes encoding
proteins that act on muscle cells to inhibit muscle cell
growth, such as myostatin, escaped negative evolu-
tionary selection. This would appear to be somewhat
of a paradox, as naturally occurring mutations in the
myostatin gene confer several benets including a
substantial increase in muscle mass in mice, dogs, cat-
tle, and even humans (13). In evolutionary terms, how-
ever, a lower muscle mass would be associated with a
reductioninbothrestingandlocomotiveenergyex-
penditure in times of food scarcity, along with the con-
servation of carbohydrate-based fuels obligatory for
preservation of brain function. Excessive muscle mass
can also lead to parturition issues (i.e., higher birth
weight and larger offspring), predisposing to evolu-
tionary disadvantages (1416). Non-muscle-related
functions of myostatin such as tendon maintenance
and repair and injury risk could have contributed to
the positive selection of this factor (17). Finally, poten-
tial trade-offs between the promotion of fatigue resist-
ance, stamina, and endurance versus muscle mass,
strength, and power could have affected the evolutionary
process (1). Accordingly, although there exists a certain
degree of synergy, distinct control and adaptation to en-
durance- and strength-based activities have evolved in
humans.
1.2. Major Themes of This Review
The importance of physical activity for health and well-
being was recognized early in human history, dating
back to records from 3000 BCE (18). The concept of
exercise is medicineand the appreciation of athletic
prowess were prominent in ancient Greek and Roman
civilizations (18). Notably, the evolutionary adaption of
humans to a phenotype eminently suitable to the pursuit
of long-distance running confers important implications
for human health and athletic performance in the pres-
ent day. Unfortunately, the fundamental link between
endurance-based activities and the evolution of numer-
ous human traits has been severely diminished in mod-
ern societies in which voluntary physical activity is at an
all-time low and has recently been exacerbated by a
global pandemic (19,20). Our twenty-rst century life-
style that in many societies encompasses round-the-
clock access to energy-dense, nutrient-poor food in the
face of prolonged periods of inactivity has resulted in
the proliferation in the rates of diagnosis of several met-
abolic disease states, a rise in morbidity and mortality,
and a high nancial burden on health care systems (21).
Paradoxically, at the same time, the standard of athletic
performance at both the amateur and professional levels
continues to advance, indicating a historically unprece-
dented divergence between the physical capabilities
of the great majority of the worlds population of inactive
individuals and a small cohort of elite athletes. Indeed,
Olympic and/or world championship medalists, world re-
cord holders, and athletes achieving within 2% of world-re-
cord performance and/or world-leading performance
comprise <0.00006% (5,000 individuals) of the entire
global population of 8 billion (22). In physiological terms,
the measure of an individualsmaximaloxygenuptake
(_
VO2max), a marker of aerobic tness, can be two- to three-
fold higher in champion endurance athletes than untrained
individuals (23). The most striking training-induced adapta-
tions contributing to such differing values are an increased
stroke volume of the heart, elevated capillary and mito-
chondrial density, and a predominance of oxidative slow-
twitchbers in the muscles of endurance-trained athletes
(24,25). While a high _
VO2max is a prerequisite for success-
ful endurance performance, this measure is also a better
predictor of morbidity and mortality than any other estab-
lished risk factor or biomarker (2630). Likewise, relative
muscle mass (3133) and strength (3436) are parameters
with high predictive power for overall morbidity and
CLINICAL HIGHLIGHTS
1) During human evolution, Homo sapiens emerged as mobile
hunters and gatherers, dependent on the natural availability of
food. However, todays sedentary lifestyle and overabundant
food availability place a major burden on our metabolic health
and are strong drivers underpinning the dramatic rise in non-
communicable diseases.
2) A sedentary lifestyle, characterized by low maximal oxygen
uptake ( _
VO2max ), unfavorable body composition, and low mus-
cle strength, is an independent risk factor for many chronic dis-
eases and a strong predictor of morbidity and mortality.
3) Despite marked interindividual differences in the response to
standardized exercise training, regular physical activity lowers
the risk of and confers therapeutic benets for many noncom-
municable diseases.
4) Endurance- and resistance-based exercise training protocols
confer distinct clinical and health-related benets and can pre-
vent or reverse many lifestyle-induced metabolic diseases.
5) Clinical exercise tests based on established, validated physio-
logical outcomes are essential for the diagnosis and subse-
quent monitoring of clinical conditions.
6) Investigations of elite human performance provide valuable
insights into the molecular, cellular, tissue, and whole body
adaptations to extreme metabolic loading. Identication of the
mechanisms and pathways that limit exercise capacity may ulti-
mately aid in the identication of novel therapeutic targets to be
prescribed to patient populations.
FURRER ET AL.
1694 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
mortality. Clearly, the biology underlying maximal en-
durance and resistance exercise performance confers
advantages beyond the athletic arena (9,37), and
while differences in physiological capacity between
elite athletes and sedentary individuals highlight the
huge disparity in performance capacity, they also pro-
vide insights into the roles of various organ systems
and the potential limits to human performance.
The last decade has seen major advances in unravel-
ing many of the putative mechanisms by which cellular,
molecular, and biochemical pathways are altered by
exercise (9,18,3841). However, many of the adapta-
tions that underpin elite athletic performance remain
poorly understood. In particular, the training programs of
world-class athletes owe more to tradition and the trial-
and-errormethods of pioneering coaches than exer-
cise biologists or sport scientists. Determining the pre-
cise role of exercise intensity, duration, and frequency in
acutely modifying various signaling cascades and coor-
dinating specic training-induced physiological adapta-
tions in athletes may offer valuable insights into some of
the critical pathways to target in order to ght the battle
against inactivity-related diseases in the general popula-
tion. Not only may sedentary or at-riskpopulations
benetfrompersonalizedphysical activity-based inter-
ventions to prevent and treat chronic lifestyle-induced
Stabilization:
-large erector spinae / gluteus maximus
-narrow waist, broad shoulders
-reduced forearm mass
-decreased facial length
-improved bio-tensegrity
Energetics:
-long, spring-like tendons
-plantar foot arch
-ankle ligaments
-long femur, short toes
-thinner heart ventricles, larger cavity
-increased hind limb muscle mass
Training
paradigms
and strategies
Nutrition
Supplements
Technological
innovations,
equipment,
facilities
Genetics
Epigenetics
Skeletal strength:
-large joint areas in lower,
but not upper limbs
Brain:
-expanded cerebello-
cerebralcortical circuitry
-cognitive abilities
-perception
-fine motor control and balance
Thermoregulation:
-eccrine sweat glands
-reduced body hair / fur
-mouth breathing
-dense skin vascularization
-large nasal epithelial area
FIGURE 1. From evolution to modern-day athleticism. Evolutionary selection of 5 main traits has facilitated the prolonged upright, bounding, bipedal
locomotion in humans. Energetic barriers are lowered by long, spring-like tendons (in particular the Achilles tendon), the longitudinal plantar foot arch,
ankle ligaments, long legs, in particular femur length, and short toes (to increase stride length and reduce vertical trajectories for better locomotor
economy), thinner heart ventricles and larger cavity, increased hind limb muscle mass, and other adaptations. Skeletal strength is conferred, e.g.,by
larger joint areas in lower but not upper limbs to dissipate impact forces. Stabilization for bipedal movement is mediated by large erector spinae and
gluteus muscles opposed to reduced forearm mass and an elongated, narrow waist and broad shoulders to facilitate counterrotation of thorax and
arms, while decreased facial length helps head stabilization or an integrated system of bio-tensegrity for embedded perturbation repelling. Eccrine
sweat glands, with a particular high density in the head for brain cooling, reduced body hair, dense skin vascularization, mouth breathing, and a large
nasal epithelial area all contribute to thermoregulation. Finally, coevolution of locomotion and the brain resulted in expanded cerebello-cerebralcortical
circuitry (for anticipation, pre-preparation, sensory integration, pre-planned multilevel compensation to deal with perturbations and destabilizations),
cognitive capabilities (to recognize landmarks, long-range orientation, recognizing prey and predators, tracking and speculative tracking/anticipation),
perception, ne motor control, and balance. See Refs. 18for more information. Modern-day athletic peak performances most likely exceed these gen-
eral evolutionary traits because of efcient training strategies and paradigms, nutrition and supplements, technological innovations, i.e., pertaining to
equipment and facilities, and genetic and epigeneticpredispositions. Figure created with BioRender.com, with permission.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1695
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
pathologies (4244), but mechanistic insights could
reveal targets for novel pharmacological interventions
(4547). A better understanding of the molecular mecha-
nisms that control skeletal muscle cell plasticity may also
provide a stimulus for further improvements in elite ath-
letic performance (48,49). The fastest 100 m sprint by a
male athlete under 18 (10.31 s, Brume Okeoghene, June
17, 2021) would have won the gold medal at the 1980
Olympics, whereas Usain Bolts 100 m world record of
9.58 s in 2009 far exceeded predictive statistical models
at that time (50). In recent years, much scienticdebate
has been focused on the limits to the mens marathon
(42.195 km) (5154).Changesinboththecultureofsport
and the recognition of modern sports science research
have supported emerging activities in which barriersto
performance have been tackled as science-driven endeav-
ors (55). The sub-2 hour marathon projectis an example:
the course design, ambient temperature, humidity, wind,
elevation above sea level, and comprehensive use of
pacemakers in highly choreographed formations helped
Kenyan runner Eliud Kipchoge run 1:59:40.2 in a specially
paced time trial in October 2019 (53,56). Likewise, there
have also been substantial advances in world-best per-
formances by female and masters-level athletes during
this time (5759).
Technological innovations in sport now drive perform-
ance enhancements at the elite/professional level, as
witnessed in track and eld (60), swimming (61), cycling
(62), and speed skating (FIGURE 2)(63), with such advan-
ces ltering down to amateur athletes, epitomized by the
widespread access to new footwear that improves running
economy (64,65). The use of novel technologies, such as
tness trackers, step counters in cell phones, or other wear-
ables, reveals behavioral aspects of physical activity linked
to performance outcomes at both a recreational and an
elite level. Such technologies can inform training design as
well as the impact of specic interventions on health and
performance outcomes (66). In the nal analysis, however,
progress in athletic performance is multifactorial, encom-
passing gene-environment interactions (6770), advances
in infrastructure, training paradigms, and design (71), nutri-
tion and ergogenic aids (72,73), as well as techniques facili-
tating recovery and regeneration, social and economic
factors, prior athletic experience and physical activity back-
ground (74), and, in an unknown number of athletes,
the use of sophisticated doping strategies (FIGURE 3)(58,
75). The range in performance capabilities, the ongoing
improvements in athletic records, and the accomplishments
of older individuals at the masters level (57)inoctogenar-
ians (76)orevencentenarians(77)alludetothevastcontin-
uum of the adaptive response of muscle tissue and other
organs to a sedentary lifestyle or exercise training. In this
review, we provide a synopsis of the training strategies of
elite athletes, the bidirectional dialogue between science,
12
11
10
9
8
7
6
5
0
1890 1900 1910 1920 1930 1940 1950 1960 1970 1980 1990 2000 2010 2020 2030
Year
Time (min)
5,000 m speed skating records
Women
Men
outdoor natural ice
outdoor artificial ice
indoor artificial ice
Improvement of ice quality Indoor 400-m rinks (1986)
Clap skate (1996)
Development of tight-fit suits
FIGURE 2. Innovations that contributed to the progress in the development of world records over time (light blue for women, dark blue for men).
Speed skating is one of many cases in which the progression of world records is driven by innovations (63). For example, the invention to improve ice
quality (natural vs articial ice) by refrigerated ovals (rst 1958), spraying tiny droplets of water to smoothen the surface (rst 1960), followed by the ice
resurfacer Zamboni(Olympics 1960) and eventually indoor rinks all contributed to new records. Additionally, the development of gear such as tight-t
suits to improve aerodynamics and the invention of the clap skate that enabled a longer contact with the ice as well as further enhanced aerodynamics
due to the crouched posture pushed the progress in world record development (http://www.speedskatingstats.com/index.php?le=records). Image on
left was taken at the 1932 Winter Olympics and is from Henriksen & Steen (public domain, via Wikimedia Commons); image on right was originally
posted to Flickr by adrian8_8.
FURRER ET AL.
1696 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
coaches, and athletes in training design and implementa-
tion, and the inherent and acquired differences between
world-class athletes and the general population (sect. 2).
Such concepts are linked to a discussion of the cellular,
morphological, and functional training-induced adaptations
in athletes (sect. 3). In sect. 4, our understanding of the mo-
lecular mechanisms that underpin the responses to acute
exercise is outlined, although these insights have largely
been obtained in non-athletes and/or animal models and
their translation to elite performance remains to be vali-
dated (78). In contrast to several previous reviews, we
address these issues for both endurance- and resistance-
based exercise training. Wherever possible, direct links
between training strategy, cellular adaptation, and molec-
ular mechanisms are discussed in an attempt to integrate
these features. Finally, we provide a discussion on
whether all individuals can become gold medal ath-
letes (sect. 5).
Gut microbiome
Education
Equipment & facilities
Sleep
Chronotype
Genetic predisposition
Epigenetic factors
Ac
Me
Me Me
Ac
Coaches
Family
Team
Knowledge
Specificity
Recovery
Intensity
Time
Lifestyle
Family history
Hallmarks
of athletes
Nutrition
Ergogenic aids
Training and recovery
Injury prevention and
rehabilitation
12:
00
Social environment
Socio-economic status
Motivation,
willpower,
perseverance,
body perception
Prior athletic
experience
N
u
r
t
u
r
e
N
a
t
u
r
e
FIGURE 3. Elite athletic performance is determined by the complex interaction of intrinsic and extrinsic factors. Undisputedly, genetic predisposition,
even though poorly dened and understood, contributes to athletic prowess and trainability. In fact, the rightgenes might even be a prerequisite for
elite, world-class performance. The epigenetic landscape is at least in part inherited but, in contrast to the genome, can also be inuenced by behavior,
including prior athletic experience, nutrition, and other lifestyle factors. A higher-than-average motivation and drive, the willpower to overcome obstacles,
adversities, and setbacks, perseverance, and the willingness to forgo activities common for non-athletic peers are essential. These factors as wellas
daily training are shaped by body perception and prior athletic experience, including a multidisciplinary/multisport practice in youths. Most likely, nutri-
tion, ergogenic aids, and gut microbiomes mutually interact in an intimate manner, collectively affecting trainability and performance. Optimal training
strategies not only comprise personalized planning but should also integrate adequate consideration of recovery and injury prevention and, if the situa-
tion arises, rehabilitation. State-of-the-art equipment and facilities are part of a permissive environment, which is also strongly shaped by socio-economic
status and social interactions with coaches, medical and other staff, team members, parents, siblings, friends, and rivals. This network of supporting peo-
ple helps to optimize knowledge and education for proper planning and implementation. Finally, peak performance also relies on proper and personal-
ized sleep patterns, matched to the individual chronotype. The use of doping might confer performance enhancements in the short term but is linked to
long-term healthdetriments and is counter to the ethos of a fair and clean sport. Figure created with BioRender.com, with permission.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1697
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
Volume
Intensity
Form
Specific
preparation Competitions Transition
Mesocycles / Phases
Macrocycle
Training focus
Periodization
HIIT
Blood flow
restriction
UndertrainingOptimal training
Functional overreachingOvertraining
Improvement
Decline
Exercise
Supercompensation
Time
Time
Progression
D
B
A
??
C
Volume
80%
Intensity
20%
75%
20%
5%
Zone 1
Zone 2 Zone 3
Polarized
Pyramidal
Performance
Functional overreaching
Performance
peak Training
Recovery
Nutrition
Psychology
Skill
General
preparation
Heat
training
Chrono-
exercise
Altitude
training
"Train low"
(glycogen)
Lactate/gas
exchange
thresholds
Critical power/
lactate turnpoint
Heterochronism of adaptations:
different systems, training paradigms, intensity,...
Performance
Fitness level
Fatigue
Homeostatic
disruption
Recovery
Repair
Regeneration
Adaptation
Involution
Adaptive
dissipation
Undertraining
Overrecovery
Optimal training
and recovery
Non-functional
overreaching
Underrecovery
Overtraining
Underrecovery
Increasing training load
Reduced recovery
Performance
Fitness level
Time
Performance
Fitness level
Time
Performance
Fitness level
Time
Performance
Fitness level
Time
Performance
Fitness level
Retrogression Taper
Recovery
Rest
Plateau
or
Maintenance
FURRER ET AL.
1698 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
2. OPTIMIZING TRAINING ADAPTATIONS TO
ENHANCE ATHLETIC PERFORMANCE
2.1. Principles of Exercise Training: Specicity,
Progressive Overload, Reversibility
A reductionist view of training for elite sport perform-
ance identies a range of interdependent adaptations
that enable an athlete to sustain the highest rate of
energy production for the duration of their event, opti-
mize economy of motion, defend cellular homeostasis,
and delay the onset of fatigue (9,54,55,79,80). In addi-
tion to undertaking workouts that promote these adap-
tations, an athlete needs to attain the optimal physique
and technical skills specic to their event(s). To achieve
these goals, elite athletes engage in periodized training
techniques involving long-term systematic planning for
major events and undertake prolonged, intense work-
outs fueled by optimal nutritional practices, while build-
ing resilience against illness and injury (55,81,82).
Coaches integrate a series of workouts that individually
target important competition performance traits into a
periodized training program composed of short (721
days) microcycles and longer (38 wk) mesocycles, cul-
minating in targeted competition peaks within a season
or year (macrocycle) (FIGURE 4). There is a rm belief
that the training-induced changes in skeletal muscle
resulting from the high-volume, high-intensity training
undertaken by elite athletes over several years is largely
responsible for the observed improvements in perform-
ance over time.
Elite athletes present a narrow range of values in many
morphological, biomechanical, physiological, metabolic,
perceptual, psychological, and other traits, depending on
their specialized event (83). Although there are multiple
and varied approaches to optimize adaptation to enhance
sporting performance based on a multitude of mechano-bi-
ological descriptors, the general principles of exercise train-
ing focus on three main concepts: progressive overload,
specicity, and reversibility. These principles of training can
be applied to individuals with a wide range of abilities
because the physiological response to specic stimuli is
largely predictable. However, the magnitude of response
of one athlete to a standardized training protocol may differ
substantially from that of another because of innate genetic
predisposition, environmental factors, access to training
facilities and sport science support, socio-cultural and eco-
nomic factors, and the interactions between these compo-
nents (FIGURE 3). The question of whether all individuals
respond to exercise training (i.e., demonstrate a measura-
ble improvement in a specied physiological outcome
measure) is discussed subsequently.
The principle of progressive overload states that once
an athlete has adapted to a given training load, the sub-
sequent training stimuli must be progressively increased
to perturb the homeostasis and thereby promote further
adaptation (FIGURE 4). Overload can be quantied
according to the volume of training (how much), the inten-
sity (how hard), and the frequency (how often), with the
magnitude of adaptation dependent on the interaction
between these variables, the prevailing tness level of
the athlete, and their genetic ceiling. In addition, cellular
and whole body homeostasis can be amplied by expo-
sure to altitude, heat, or altered fuel availability (55). Such
approaches are based on the premise that by imposing
greater metabolic stressand provoking extreme distur-
bances to homeostasis, intracellular responses in skeletal
muscle (and other tissues and organs) will be maximized,
thereby invoking superior training adaptation and
enhancing one (or more) of the factors underpinning per-
formance (84). Several training strategies are currently
practiced by competitive athletes in the belief that they
amplify adaptation and lead to improved performance
capabilities. Here, we describe a selection of training
strategies that have high uptake by elite athletes and
have plausible biological mechanisms that might explain
current practices (8587).
The principle of specicity states that any training-
induced adaptations that accrue to an athlete are unique
to the type of exercise mode performed; this is most evi-
dent when contrasting the divergent phenotypes that
result after undertaking either prolonged endurance- or
FIGURE 4. Training principles and strategies. A: to achieve peak performance at the timeof competition, training volume, intensity, and form/specic-
ity have to be adapted in different cycles/phases. Specic paradigms, e.g., high-intensity interval training (HIIT), train low,and others, are likewise
periodized and matched to the prevailing volume/intensity/form requirements. Importantly, the periodization of training has to be matched to that of
nutrition (e.g., low glucose vs. carb loading), recovery, psychological aspects, and skill acquisition. B: within shorter cycles, e.g., weekly planning, polar-
ized or pyramidal partitioning of training volume at different intensities (e.g., dened by lactate/ventilatory thresholds between zones 1 and 2and the
critical power/lactate turnpoint between zones 2 and 3) helps to improve performance and reduce overtraining. C: training adaptation is initiated by a
homeostatic disruptioninduced by exercise. After exercise cessation, recovery and repairmechanisms not only result ina return to baseline but trigger
adaptive mechanisms, optimally in a supercompensatory manner, which should help to protect muscle better from future perturbations. However, in
the absence of continued stimuli, i.e., detraining, this supercompensatory response is abolished by an adaptive dissipation. The amplitude and tempo-
ral aspects of this curve are strongly inuenced by the training paradigm and related parameters. Moreover,within the same system, biochemical proc-
esses, cell types, or tissues can react in a different manner (heterochronism of adaptation). D: performance gains are controlled by the balance
between training load and recovery. A suboptimal planning can result in either undertraining with little or no gains or overtraining, in which performance
decreases (retrogression) and the risk for injuries increases. In proper conditions, a functional overreach helps to maximize progression and overcome
training plateaus. Figure created with BioRender.com, with permission.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1699
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
strength/resistance-based training (86,88). The principle
of specicity states that the closer the training routine is
to the requirements of competition, the greater the likeli-
hood of successful outcomes. For this reason, the foun-
dation of any training program should reect the desired
training adaptation necessary to enhance sports-specic
performance. The principle of specicity should operate
with regard to not only the modality of training but also
the intensity and speed/power output at which an athlete
performs training (discussed below). The principle of re-
versibility states that there will be a decline or complete
loss of training-induced adaptations when an athlete
reduces or stops training for a substantial time (i.e., sev-
eral weeks up to several months). Reductions in both
training volume and intensity diminish many of the adap-
tations that accrue from daily or twice-a-day training, with
such a response leading to concomitant performance
decrements. The time courses of loss of adaptations af-
ter both well-trained endurance athletes and recreational
sportspersons stop daily training are rapid: declines in
whole body maximal and submaximal responses to exer-
cise occur during the rst 721 days of inactivity, becom-
ing somewhat stable after 2 mo of detraining (8993). In
athletes who predominantly train for strength and power,
and depending on the type of strength test performed,
there is a limited decline in muscular strength during
short-term (up to 21 days) inactivity, but decay rates
increase substantially after 4 wk and longer (89,9395).
It is important to highlight that the principle of reversibility
differs from a competitive taperbefore a major event/
competition: during a taper, the volume and frequency of
training are deliberately reduced but the intensity is
maintained or even increased, resulting in a performance
enhancement of 12% (96,97).
2.2. Intensity vs. Volume to Optimize Training
Adaptation
2.2.1. High-intensity, low-volume vs. low-intensity,
high-volume training to maximize endurance
training adaptation.
Recently, there has been renewed scienticinquiryalong
with widespread public interest in the potential for high-in-
tensity interval training (HIIT) to induce physiological adap-
tations that are similar or even superior to a traditional,
continuous endurance-based exercise prescription for
health and performance (98100). HIIT broadly refers to
exercise that is characterized by relatively short bursts of
vigorous activity interspersed by periods of rest or low-in-
tensity exercise for recovery. A common classication sub-
divides this type of training into 1) sprint interval training
[SIT, supramaximal efforts performed at power outputs/
speeds >peak oxygen uptake ( _
VO2peak), for 3060 s,
with 1- to 3-min rest or active recovery], 2) high-intensity
interval training (HIIT, comprising near-maximal efforts per-
formed at the power output/speed that elicits _
VO2peak for
14 min, with 1- to 2-min rest or active recovery), and 3)
moderate-intensity continuous interval training (performed
at power output/speed that elicits between 85% and 90%
of _
VO2peak performed for 510 min, with 1-min rest or
active recovery). In untrained and recreationally active
individuals, both short-term SIT and HIIT are potent stimuli
to induce physiological remodeling similar to that attained
after traditional prolonged endurance training, despite
markedly lower total exercise volume and training time
commitment (101,102).
The notion that interval training is a new, groundbreak-
ing scientic approach to physical conditioning, espe-
cially for athletic performance, needs to be placed in
historical context. Coaches and athletes have appreci-
ated the value of this form of training since the early twen-
tieth century, with many notable cases in which a range
of different work to rest intervals were trialed, tested, and
rened to prepare for competition (99). Interval training
was widely used by a Finnish coach, Lauri Pikhala, who
worked with many champion runners including Paavo
Nurmi and Hannes Kolehmainen. Between 1920 and
1930, Nurmi was the most dominant distance runner in
the world, winning a total of nine Olympic gold medals.
The foundation of Pikhalas training methods focused on
running a high number of repetitions (2030 efforts) at
close to the athletes race pace interspersed with short
(<60 s) rest intervals. Subsequently a German physician
and coach, Woldemar Gerschler, working with cardiolo-
gist Herbert Reindel, ne-tuned a similar interval training
approach focusing on the manipulation of the work:recov-
ery periods, based on an athletes heart rate. An athlete
would run over a distance fast enough to elicit a heart
rate close to 180 beats/min, after which they rested until
the heart rate dropped to 120 beats/min; at this time,
the next work bout was performed. Gerschler and
Reindel proposed that the rest or recovery interval was
the most important aspect of their approach because it
was during this phase that the heart adapted, allowing it
to grow larger and stronger (99). In the 1960s, the New
Zealand running coach Arthur Lydiard advocated a shift
away from high-intensity interval-based training to high-
volume, continuous training for endurance performance.
Lydiardadvocatedrunningasmuchas160km/wkduring
the preseason conditioning or basephase, with both
middle- and long-distance runners undertaking similar
volumes of work (103). Although there was a perception
that such a high volume of training could only be per-
formed at low intensities (i.e., high volume, low intensity),
this was not the case: running during this phase of condi-
tioning was prescribed at speeds that corresponded to
an athletes best 16 km race pace (for middle-distance
FURRER ET AL.
1700 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
athletes) or best marathon pace (for long-distance run-
ners). This conditioning phase could last from as short as
8 wk to half a year. Lydiardsathleteshadmajorsuccess
over two Olympiads (Rome 1960 and Tokyo 1964), win-
ning medals across a wide range of distances including
triple Olympic gold medalist Peter Snell (800 and 1,500
m), John Davies (bronze medal 1,500 m), Murray Halberg
(gold medal 5,000 m), and Barry Magee (bronze medal,
marathon).
Despite these successful coach-driven approaches to
conditioning for elite athletes, it was not until the 1960s
that the rst scientic publications on the physiological
bases of training for human performance appeared, and
even today the scientic literature on the unique effects of
specic training interventions on the performance of
highly trained athletes is sparse. Indeed, although the
foundation of all training programs for the enhancement of
sport performance is the volume, intensity, and frequency
of exercise, the relative importance of these interdepend-
ent variables has not been established for many of the
key physiological adaptations to training, nor their impact
on performance outcomes (104,105). This is because train-
ing prescription is innitely variable, with countless permu-
tations around the core tenets of the general principles of
training (FIGURE 4). Adding to the complexity of training
prescription is the multiplicity of the physiological/technical
demands of many athletic events, with many requiring
components of both endurance and strength/power, as
well as different modes of exercise (i.e., swimming, cycling,
and running in the triathlon). Potential interference
effectsbetween endurance- and strength/power-based
training regimens are discussed below.
There has been spirited scientic debate as to whether
training volume or training intensity promotes the greatest
adaptation in skeletal muscle (104,105), with this dialogue
focusing predominantly on exercise-induced changes in
mitochondrial content, typically assessed by quantifying
the maximal activity of citrate synthase, the rst step of the
tricarboxylic acid cycle, or skeletal muscle respiratory
capacity (see sects. 3.4.1 and 4.5). Although higher inten-
sities of exercise generally elicit greater increases in mito-
chondrial content than lower exercise intensities per unit
of time or work (104), such a narrow perspective ignores
any functional outcomes, such as athletic performance.
Perhaps more to the point, the data used to support one
or the other position (i.e., volume vs. intensity overload)
have come from studies that employed untrained or rec-
reationally active subjects participating in short-term inter-
ventions (26 wk) undertaking one-dimensional training
programs consisting of either HIT or continuous, submaxi-
mal endurance-based training. It is not clear how these
results can be extrapolated to elite athletes with a pro-
longed history of periodized training that includes a vari-
ety of workouts with different goals, performed within
well-dened training cycles, at volumes, frequencies, and
absolute exercise intensities/power outputs that far
exceed those capable of being attained by their less ge-
netically gifted counterparts. In this regard, a recent study
reported reductions in mitochondrial respiration in skeletal
muscle in response to 4 wk of intensied HIT in moder-
ately trained individuals (106). The impairment in mito-
chondrial function occurred during the week of heaviest
training load but was dissociated from both mitochondrial
activity and mitochondrial protein abundance, which both
peaked at that time (106). Despite the transient impair-
ments to mitochondrial respiration, performance parame-
ters all increased after the intensied HIT regimen.
Furthermore, the training undertaken by the participants
in that study consisted exclusively of maximal HIT (106)
and can only be tolerated by highly trained athletes for
more than a few successive days (107).
Since the classic model of training periodization was
rst proposed over four decades ago (108), there has
been widespread discussion about how best to imple-
ment training stimuli to optimize adaptation and athletic
performance (109). Although several long-term period-
ization approaches have been described (110), con-
trolled studies comparing the impact of these different
protocols on performance outcomes are lacking. As
noted, anecdotal testimonies from top athletes and their
coaches (111), case studies of elite performers (112,113),
and reports of small cohorts of top athletes from specic
sports (53,114,115) provide insights into the training
practices of elite performers, but such studies merely
document what successful athletes did; they do not
reveal what made those athletes successful or prove
that the program they followed was optimal (116).
Indeed, there may have been many athletes who fol-
lowed similar programs who were not successful, fell ill,
suffered injury, or dropped out of the sport completely.
Notwithstanding these limitations, detailed analyses of
the training methods of elite athletes enable sport scien-
tists to examine relationships between training inputs
and variables directly or indirectly related to perform-
ance (FIGURE 5). This information can also provide a ba-
sis for hypothesis testing with respect to training load
and physiological adaptation. There have been multiple
empirical descriptions of the distribution of training in-
tensity in highly trained/elite athletes competing in en-
durance-based sports (110,114,117121). Depending on
the specic loading characteristics of the sport (i.e.,
weight bearing vs. non-weight bearing), international
athletes competing in endurance events typically train
for between 500600 h (distance running) and up to
1,000 h per year (rowing, swimming, cycling, triathlon),
performed during 400800 training sessions (122124).
This training volume is undertaken for a minimum of 11
mo a year, with the overall goal of achieving peak
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1701
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
performance throughout a specied time frame (usually
46 wk) in the competitive season. However, there is
signicant variability between sports, with professional
cyclists frequently racing 100 days and riding in excess
of 30,000 km during any 12-mo period (125,126).
Longitudinal data suggest that the development of a
world-class endurance athlete may take up to a decade
of specic training, with highly successful athletes often
following a 2- or 4-yr cycle of preparation for world
championships or Olympic events (110,125). To maxi-
mize adaptation and reach ones genetic potential,
champion athletes must therefore be able to tolerate
prodigious training loads. However, a high training vol-
ume alone does not guarantee sporting success.
A quarter century ago, Mujika et al. (127)studiedthe
relationships between training variables and performance
variations over the season in a group of elite swimmers.
They reported that training intensity, rather than volume
or frequency, was the key variable inducing a training ad-
aptation that led to subsequent performance improve-
ments. These workers also observed a training intensity
distribution that placed emphasis on volume-overload
training conducted at submaximal intensities for most of a
season, with the inclusion of supramaximal high-intensity
sprint workouts nearer to a competition. A decade later,
this approach would be described as a polarized training
intensity distributionby Seiler and colleagues (121)asdis-
tinct from a pyramidal training paradigm (FIGURE 4).
Since then, there have been several reports that elite ath-
letes follow both approaches to their competition prepa-
ration (128). Coetzer et al. (129) reported that elite
distance runners with superior race performances trained
at a higher average intensity than a group of sub-elite run-
ners: the sub-elites spent 13% of their total weekly training
volume running at speeds eliciting >80% of _
VO2peak,
whereas the elite runners spent signicantly more time
(36%) at this higher intensity. These observations agree
with others (118,119) who have observed that elite Kenyan
distance runners complete a greater volume of training
as fast-paced temporuns and short-interval training
compared to their non-elite counterparts. Guellich and
colleagues (130) reported that elite endurance ath-
letes from a range of sports including rowing, running,
cycling, and cross-country skiing perform only a small
portion of their training at competition/race-pace
intensities, with the bulk of their workload comprising
low-intensity, high-volume workouts and exposure to
extreme HIT sessions.
It has been hypothesized that a polarized approach to
training, in which 7580% of total training volume is
FIGURE 5. Periodization of training for an elite athlete. Schematic representation of periodization of training along with physiological data collected
during preparation for the 2016 Rio Olympic Games for a gold medal-winning female rower. BLa, blood lactate; BM, body mass. See GLOSSARY for other
abbreviations. Figure created with BioRender.com, with permission.
FURRER ET AL.
1702 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
performed at low intensities with 1015% performed at
supramaximal intensities, may be the optimal training in-
tensity distribution for elite athletes who compete in
intense endurance-based events (131). However, this prac-
tice has recently been questioned and debated (116,132
134). Alternative approaches to polarized traininghave
been proposed, such as pyramidal or thresholdtraining
intensity distributions. At present, and to the best of our
knowledge, there are no studies that demonstrate that
adherence to a polarized training program produces
superior outcomes compared with the pyramidal training
programs athletes typically practice or other possible
training models (116,132134). Indeed, polarized training
per se seems totally incompatible with the principle of
training specicity, a cornerstone of any training program.
Although it is tempting to attribute the superior perform-
ances of elite athletes from a range of endurance sports
to the adoption of a specic training regimen (i.e., polar-
ized training, HIT), the principle of individuality dictates
that the same training program will not equally benetall
those who undertake it. Furthermore, the molecular and
cellular mechanisms that underpin performance enhance-
ment after polarized and various other training interven-
tions are not well understood. Directly linking exercise-
induced molecular signaling events in skeletal muscle to
dened metabolic responses and specic changes in
gene and protein expression that occur after diverse train-
ing regimens may provide clues as to why certain training
methods (i.e., polarized training, HIT) are such potent
interventions both for promoting health outcomes and
enhancing athletic performance.
2.2.2. High-intensity, low-volume vs. low-intensity,
high-volume training to optimize resistance
training adaptation.
Analogous to endurance-based training, periodization is
frequently used to promote muscle hypertrophy and
strength gains in response to a program of resistance
training (135). Indeed, when resistance training volume is
similar, periodized training protocols induce greater
gains in strength [i.e., one-repetition maximum (1RM)]
than non-periodized resistance training, at least in
trained individuals (135). The process of skeletal muscle
ber hypertrophy and the concomitant gains in strength/
power (discussed in sects. 3.2, 4.34.5, and 4.8) are the
result of the conuence of a net positive muscle protein
balance, with the addition of satellite cells to muscle
bers a possible mechanism. Muscle hypertrophy only
occurs when net positive muscle protein balance is
maintained over several weeks/months and when the
rate of muscle protein synthesis (MPS) exceeds that of
muscle protein breakdown (136). Resistance training vol-
ume can be dened as the number of sets repetitions,
sets repetitions load (expressed as a percentage of
1RM), sets repetitions load (kg), load sets repeti-
tions for each exercise, or number of sessions repeti-
tions sets (137). Resistance training intensity is typically
dened as a percentage of maximal strength (%1RM).
Resistance training frequency represents the number of
resistance training sessions performed in a specied
time period (i.e., per week) and for each muscle group.
The frequency of resistance training sessions is impor-
tant when considering resistance exercise prescription,
as the recovery time between sessions must allow for
muscle adaptation (i.e., net protein synthesis). The num-
ber of training sessions provides an indication of the total
resistance training work over a programsduration,
whereas including load describes the total work of a sin-
gle training session. Other parameters (e.g., load, num-
ber of repetitions and sets, range of movement, time
between sets, time under tension, and volitional muscle
failure) provide a comprehensive description of resist-
ance training programs (138), even though the effect of
manipulating these variables on athletic performance
remains unclear (139).
To maximize muscle hypertrophy, the American
College of Sports Medicine (ACSM) recommends resist-
ance training intensities corresponding to a load of
7080% 1RM for 812 repetitions (140). Although such
loading is unlikely to be undertaken by elite athletes
competing in strength/power events, such advice is
largely based on the observation that higher loading
induces greater force development, an increased mus-
cle electromyography activity (141), and a greater recruit-
ment of muscle bers. Evidence to support a dose-
response relationship between external loading and
maximal rates of MPS comes from the work of Kumar et
al. (142), who showed that a plateau in MPS was reached
at intensities approaching 90% of 1RM. However, results
from other studies suggest that maximal rates of MPS
can also be achieved by low-intensity, higher-volume
loading. Burd et al. (143) studied 15 recreationally active
males who performed four sets of unilateral leg exten-
sion exercise at different exercise loads and/or volumes:
90% of 1RM until volitional failure (90FAIL, 5repeti-
tions), 30% 1RM work-matched (WM) to 90%FAIL (30WM,
14 repetitions), or 30% 1RM performed until volitional fail-
ure (30FAIL, 24 repetitions). Low-load, high-volume re-
sistance exercise (30FAIL) was equally effective at
increasing rates of MPS as high-load, low-volume resist-
ance exercise (90FAIL), eliciting increases in rates of
myobrillar protein synthesis similar to those induced by
the 90FAIL protocol in the postexercise recovery period.
Furthermore, only the 30FAIL protocol sustained higher
rates of MPS 24 h after exercise. Although these data
from a single bout of resistance training are intriguing,
there is support for the concept that measures of acute
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1703
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
postexercise MPS are qualitatively predictive of the
chronic training-induced phenotypic changes driven by
repeated resistance exercise stimuli. In a study from the
same laboratory, Mitchell et al. (144) studied 18 untrained
males who completed 10 wk of unilateral knee extension
resistance training. Each leg of a participant was randomly
assigned in counterbalanced fashion to one of three possi-
ble unilateral training conditions: one set of knee extension
performed to voluntary failure at 80% of 1RM (80%-1); three
sets of knee extension performed to the point of fatigue at
80% of 1RM (80%-3); or three sets performed to the point
of fatigue with 30% of 1RM (30%-3). Each participant
trained both legs and was therefore assigned to two of
the three possible training conditions. The strength of
this design is that both limbs are exposed to the same
nutrient and hormonal milieu and therefore any pheno-
typic changes can be ascribed solely to the training stim-
ulus. There were signicant training-induced increases
in muscle volume [measured by magnetic resonance
imaging (MRI)], but these were not different between the
training protocols. These results are in accordance with
previous acute measurements of muscle protein syn-
thetic rates and demonstrate that a lower load lifted to
failure results in muscle hypertrophy similar to a heavy
load lifted to failure. An important feature of this study
was that the training program was underpinned by
adequate nutrition (i.e., sufcient amino acid availability)
to support the increases in MPS that occur after each
training session. These results support earlier ndings
demonstrating that signicant increases in muscle ber
area can be achieved after 16 wk of isometric training at
30% of maximal voluntary contraction (MVC) (145).
As is the case for most studies that have examined
various endurance training protocols, most investiga-
tions of different strength/resistance training programs
have been undertaken with recreational and/or moder-
ately trained male college students. How such ndings
translate to elite athletes who are likely to have reached
an upper limit in muscle hypertrophy and strength gains
after many years of training is unclear. Elite athletes
competing in events that require strength/power will
also be undertaking additional forms of training to maxi-
mize muscular force output such as plyometrics, which
involves rapid and repeated stretch/contractions of the
muscle of the lower limb (146148), or hypoxic/blood
ow-restricted training (149151). This makes it difcult to
determine the precise contribution of any single inter-
vention to improvements in muscle hypertrophy and
strength. Inherent variability in the individual response
to resistance training is also a factor to consider in any
training protocol (discussed below). In summary, there is
currently little consensus on how the variables related to
resistance training (training load, volume, and fre-
quency, muscle time under tension, lifting cadence,
contraction mode, and interset rest interval) are most
effectively periodized to maximize both MPS and
improvements in strength and other functional meas-
ures (135,139,152,153).
2.3. Exercise Interference Eects and Concurrent
Training Responses
The inverse relationship between muscle ber size and
oxidative capacity highlights the principle of the specic-
ity of training when comparing muscles of endurance
and strength/power athletes (154,155). Accordingly,
simultaneously training for both endurance and strength
results in a compromised adaptation compared with
training for either exercise modality alone, at least in pre-
viously untrained individuals. This phenomenon was rst
described by Hickson (156), who reported impaired
strength development in training naive males when they
incorporated both strength and endurance workouts
versus single-mode exercise into a short-term (10 wk)
training program. Hickson (1980) coined this the inter-
ference effect,and since that seminal observation, a
number of animal and human studies have been con-
ducted in an effort to elucidate a molecular basis to
explain this outcome (discussed in sect. 4.5.1). Of note
was that training-induced gains in aerobic capacity in
that study (156) were not compromised by concurrent
strength and endurance training. In fact, in contrast to
the impaired strength gains observed when endurance
training is undertaken simultaneously with resistance
training (156), there is potential for combined strength
and endurance training to amplify endurance perform-
ance (157).
The study of concurrent traininghas received less
scientic enquiry than single-mode training for endur-
ance or strength/power. Indeed, studies of concurrent
training interventions pose several unique experimental
challenges. The inability to match total work as well as
the type of stimulus and/or exercise mode makes com-
parisons between the results of studies of concurrent
training problematic. Differences in experimental design
and dependent variable selection also limit any mecha-
nistic insights in those studies that have determined
only performance-based outcomes. Finally, the majority
of studies of concurrent training to date have focused
exclusively on acute molecular responses in moderately
trained individuals, employing modest workloads; the
training practices of elite/professional athletes undertak-
ing concurrent training far exceed those reported in the
literature for less well-trained subjects and are likely to
induce complex molecular proles (88,158). Over the
past two decades, the mechanisms that generate the
adaptive response to both endurance- and strength-
based exercise training have undergone intense
FURRER ET AL.
1704 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
investigation (9,55,8587,159166). There are multi-
ple stimuli associated with endurance- and resistance-
based exercise and various signaling kinases that
respond to these different perturbations, in concert
with numerous downstream pathways and targets of
these kinases. These events involve the increased
expression and/or activity of key proteins mediated by
an array of signaling events, pre- and posttranscrip-
tional processes, regulation of translation and protein
expression, and modulation of protein/enzyme activ-
ities and intracellular localization (9,55,85,87,159).
These molecular processes are described in detail in
sect. 4. Finally, there are complex spatial and temporal
interactions between the various elements that ulti-
mately combine to produce the integrated response
to an exercise challenge that, when repeated over
months and years, results in functional improvements
in performance and alterations in phenotype.
Although it is convenient to classify athletic events
as either endurance-basedor strength-/power-based,
with skeletal muscle from endurance- and strength-
trained individuals representing diverse adaptive
states in response to selective activation and/or
repression of signaling pathways that underpin these
adaptations (9,88,159,160,166), such a one-dimen-
sional perspective ignores the fact that the majority of
athletic disciplines require a combination of both mus-
cular endurance and strength/power for successful
outcomes. As such, both endurance- and resistance-
based training are frequently undertaken concomi-
tantly as part of a periodized training program. These
practices encompass several scenarios: 1)asingle
training session during which an athlete performs
both endurance- and resistance-based exercise; 2)
two independent training sessions undertaken by the
athlete on the same day, in one of which the focus is
endurance adaptation (i.e., performed in the morning)
and in the other strength/power adaptation (i.e., per-
formed in the afternoon/evening); or 3)whenanath-
lete incorporates both types of training on different/
alternate days as part of a periodized training program
(88). Currently, little is known about the effects of con-
current training in elite athletes on performance pro-
gression, and it is conceivable that the degree of
interference may be discipline- and training paradigm-
specic(1,167). For example, in sports where endur-
ance as well as high peak power/forces are required
(such as in 2,000 m rowing), athletes aim to maximize
both muscle mass and oxidative capacity. Indeed, the
peak power of Olympic rowers is positively correlated
with thigh muscle volume but negatively correlated
with _
VO2max (168). Similarly, sprint and endurance per-
formance are inversely related in highly trained
cyclists (169).
2.4. Altitude and Hypoxic Training to Enhance
Adaptation
Of all the practices currently used to enhance training
adaptation and elite athletic performance, altitude train-
ingor exposure to hypoxic environments is the most
widespread (55,97). The stimulus for a new era in
research of high-altitude training practices was the 1968
Olympic Games held in Mexico City at an elevation of
2,240 m above sea level. In the middle- and long-dis-
tance track events, runners who were born and trained
at altitude were dominant: in the mens10,000m,the
rst ve runners resided and trained at altitude. The
world record holder at the time for both the 5,000 and
10,000 m events going into the Mexico Games,
Australian Ron Clarke, who was born and trained at sea
level, collapsed after nishing 6th in the 10,000 m and
had to be administered oxygen to recover. Since those
Olympics, male and female athletes from Kenya and
Ethiopia have dominated middle- and long-distance run-
ning events, with elite athletes and coaches steadfastly
believing in the benets of hypoxia-induced adaptive
responses to optimize performance (97). This is despite
the paucity of scienticevidencesupportinganaltitude-
induced performance-enhancing effect (55,170,171).
The mechanisms that underpin the adaptive response
to reduced oxygen availability are discussed below.
2.4.1. Into thin air: altitude training strategies to
enhance endurance performance.
There are several common approaches that athletes
adopt with regard to altitude training, involving several
days to several weeks of exposure to some form of alti-
tude or hypoxic challenge (172). Regardless of the differ-
ent approaches used to induce hypoxic living/training
conditions, the underlying physiological basis for alti-
tude training is that the reduced barometric and partial
pressure of oxygen results in lowered oxygen availabil-
ity causing an increase in erythropoietin (EPO) produc-
tion in the kidney that stimulates erythropoiesis and
thereby leads to enhanced hemoglobin (Hb) mass. As
acute exposure to hypoxia over several hours does not
improve aerobic or anaerobic performance, these studies
are not discussed here (173). The original altitude training
strategy involved athletes spending up to 6 wk living and
training at a moderate altitude (2,0002,500 m) and
returning to sea level just before a major sea-level com-
petition (live high, train high,LHTH). The LHTH
approach boosts EPO and Hb mass, which results in an
increase in _
VO2max. Such adaptations usually persist for
12 wk upon return to sea level, with the athlete partici-
pating in several major competitions during this period. A
limitation of the LHTH strategy is that training intensity is
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1705
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
often compromised, which is in line with the linear reduc-
tion in _
VO2max with increasing altitude (68% reduction
per 1,000 m) (174). A second strategy involves athletes
continuing to reside at sea level but training at altitude
(live low, train high,LLTH). Adaptations resulting from
LLTH are mainly conned to the trained musculature (i.e.,
skeletal muscle mitochondrial volume density), with little
effect on EPO or Hb mass. As with the LHTH approach,
the intensity of training is typically reduced with LLTH. A
third protocol, and the one that is most widely used and
has received widespread interest among scientists,
coaches, and athletes, is the live high, train low(LHTL)
approach, whereby athletes reside at altitude for several
weeks but return to sea level to undertake the majority of
their training sessions. Compared with LHTH or LLTH
approaches, the LHTL approach permits athletes to main-
tain their absolute training loads (volume and intensity)
while concomitantly gaining the physiological adaptations
that accrue with exposure to chronic hypoxia. Indeed,
when competitive runners completed 4 wk of supervised
training as either LHTL, LHTH, or LLTH, performance of a
5kmtimetrialatsealevelwasimprovedonlyintheLHTL
athletes despite similar gains in the athletes_
VO2max in all
intervention groups (175). No muscle biopsies were taken
in that investigation, so it was not possible to determine
whether the different altitude-training regimens resulted
in changes in hypoxia-mediated signaling pathways or if
there were changes in major training-induced signaling
proteins. A model pioneered by the Australian Institute of
Sport (AIS) requires that athletes gain exposure to alti-
tude/hypoxia by either living in a custom-built altitude
house under conditions of simulated altitude (14 h/day) or
using altitude tents or intermittent hypoxic exposure with
hypoxic breathing devices (176). However, even though
altitude paradigms increase Hb mass (172), the purported
performance gains from living at simulated moderate alti-
tude and training at low altitude have been questioned
(177,178). Therefore, whether training in hypoxia while liv-
ing in normoxia or living under hypoxic conditions while
training at sea level (or low altitudes) is superior to living
and training in normoxia for enhancing performance of
elite athletes near sea level is unclear and warrants fur-
ther investigation. There are also many challenges when
assessing the effect of altitude exposure on performance
in elite athletes (179). For example, the scienticgold
standard design of a double-blind, placebo-controlled,
crossover trial has seldom been conducted in studies of
altitude training in elite athletes. A recent systematic
review, albeit incorporating individuals with a wide range
of athletic abilities, concluded that placebo and nocebo
effects can exert a small to moderate effect on sports per-
formance (180). Yet despite equivocal scienticevidence
to support a performance-enhancing effect of altitude/
hypoxic training practices, elite endurance athletes
and their coaches continue to believe that some form
of altitude training will confer a performance advant-
age when competing at sea level. Guidelines and
measures to improve altitude acclimatization, toler-
ance, and safety have been reviewed elsewhere (181).
Interestingly, preconditioning with hyperbaric oxygen
has also been proposed to enhance performance,
however with similar equivocal underpinnings (182).
2.4.2. Resistance training under hypoxic
conditions.
Acute hypoxia has been proposed to potentiate resist-
ance training-induced hypertrophy by activating satellite
cell-dependent myogenesis rather than an improvement
in net protein balance. To test this hypothesis, van
Doorslaer et al. (183) recruited 19 physically active male
subjects who performed 4 wk of resistance training (6
sets of 10 repetitions of a 1-leg knee extension exercise
at 80% 1RM 3 times/wk) in either normoxic [fraction of
inspired oxygen (FIO2): 21%; n= 9] or hypoxic (FIO2:13.5%,
n= 10) conditions. At the end of the intervention, the
strength gain was higher in individuals who trained
under hypoxic compared with normoxic conditions, de-
spite no changes in muscle thickness and the rate of
MPS. Although these results suggest that training under
hypoxic conditions may be a potent intervention to
increase muscle strength, at least in the early phase of
training, additional studies in well-trained athletes incor-
porating long-term protocols are urgently needed to
determine whether hypoxic resistance training can fur-
ther maximize strength gains. Other protocols with
potential additive training effects due to reduced local
muscle oxygen availability and exacerbated vascular
shear stress that leverage hypoxic stimuli (i.e., blood
ow restriction) are currently being investigated (150,
151,184) yet hampered by the heterogeneous responses
to ischemic preconditioning (185).
2.5. The Lowdown on Training with Reduced
Muscle Glycogen Stores
Agrowingeld of interest that has directly risen from a
better understanding of the molecular bases underlying
training adaptation is how nutrient availability has the
capacity to modify the regulation of many contraction-
induced signaling networks in skeletal muscle (sects.
3.4.2 and 4.5) (9,186193). The interaction between exer-
cise training-induced responses and nutrient availability
has long been recognized (194), and today it is well
accepted that carbohydrate-based fuels are critical for
prolonged, intense training and in the competition setting
where optimal endurance performance is desired (195).
However, this premise does not address the issue of
FURRER ET AL.
1706 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
whether training adaptation is driven by a surplus or lack
of substrate (i.e., carbohydrate). During the past decade,
there has been a growing appreciation that commencing
selected training sessions with reduced muscle glycogen
stores may promote training adaptation and enhance en-
durance performance (196198). Acutely manipulating
substrate availability (by either altering the composition
and/or timing of meals before training/competition or
depleting endogenous fuel stores by exercise) rapidly
alters the concentration of blood-circulating substrates
and hormones that modulate several receptor-mediated
signaling pathways. The release of cytokines and growth
factors from contracting skeletal muscle in response to
the altered hormonal milieu also stimulates cell surface
receptors and activates many intracellular signaling cas-
cades (described in sect. 4). These local and systemic fac-
tors cause marked perturbations in the storage prole of
skeletal muscle (and other insulin-sensitive tissues) that,
in turn, exert pronounced effects on resting fuel metabo-
lism and patterns of fuel utilization during exercise. When
repeated over weeks and months, such nutrient-exercise
interactions have the potential to alter numerous adaptive
processes in skeletal muscle that ultimately drive the phe-
notype-specic variability observed between individuals
(55). However, linking these molecular events to direct
downstream effectors has proven elusive (199). Perhaps
more to the point, training adaptation requires an
increase in the steady-state levels of exercise-induced
proteins, and it was not until the pioneering study of
Hansen and colleagues (187) that the notion that endur-
ance training undertaken with low muscle glycogen lev-
els could augment adaptation gained scientic credibility.
These workers tested previously untrained individuals
before and after a 10-wk intervention in which both the
left and right legs of the same individual were subjected
to specic work-matched training protocols in which one
leg was trained once daily while the contralateral limb
trained twice every second day. As intended, the twice-a-
day training protocol decreased muscle glycogen content
after the rstboutofexercisesuchthatthesecondexer-
cise session of the day was commenced with lowered
(but not totally depleted) muscle glycogen content. The
activity of mitochondrial enzymes along with resting mus-
cle glycogen concentration were all increased to a
greater extent when half the training sessions were exe-
cuted with low glycogen availability. Exercise time to
exhaustion (a proxy for performance) involving a one-leg-
ged kickingtask was elevated markedly for both legs
after training but was twice as long for the limb that
trained with low compared to high glycogen. The
strength of this study was the design that controlled for
both systemic and local effects. However, the authors
acknowledge that the controlled laboratory setting,
coupled with the training status of their subjects, may not
permit the results to be extrapolated to competitive ath-
letes. Several studies subsequently veried the nding
that, in well-trained athletes, chronic (310 wk) training
programs in which selected workouts were deliberately
commenced with low muscle glycogen concentration
increased the expression of genes and the abundance of
proteins involved in carbohydrate and/or lipid metabolism
while promoting mitochondrial biogenesis to a greater
extent than when all workouts are undertaken with nor-
mal or elevated glycogen stores (molecular mechanisms
underlying these observations are discussed in sect.
4.5.2) (199201). Surprisingly, such adaptations accrued
notwithstanding a reduction of 78% in the athletesself-
selected training intensity (200,201). Yet despite aug-
mented adaptations at the muscle level, studies that have
examined the train lowglycogen model in well-trained
athletes have often (201204), but not always (197,198),
failedtoshowaperformancebenet(forreview,seeRef.
205). Such a disconnect between changes in selected
molecular mechanistic variables (e.g., increases in the
phosphorylation status of signaling molecules and/or
increases in the expression of genes and proteins
involved in mitochondrial biogenesis) and whole body
functional outcomes (changes in training capacity or ath-
letic performance) is hard to reconcile. However, it may
well be that elite athletes with a prolonged history of train-
ing have already maximized many of the cellular path-
ways involved in energy provision and that proteins in
these and other contraction-induced pathways that are
upregulated with the train low glycogen protocol are not
rate limiting for performance.
There is a scarcity of studies that have examined the
effects of commencing resistance training with low mus-
cle glycogen stores. Nevertheless, some evidence
exists suggesting that reduced glycogen availability may
upregulate cellular pathways regulating mitochondrial
biogenesis after a single bout of exercise (206), even
though engaging in resistance training with low muscle
glycogen does not affect rates of MPS (207). These
results imply that commencing a bout of strenuous re-
sistance exercise with low muscle glycogen availability
attenuates neither anabolic signaling nor rates of myo-
brillar protein synthesis. In summary, despite no clear
evidence of a performance-enhancing effect from the
results of several well-controlled laboratory-based stud-
ies that have tested various train low (glycogen) strat-
egies, many athletes who compete in endurance-based
events continue to incorporate such practices into their
training programs. In contrast, there appears no reason
for athletes undertaking resistance training regimens to
adopt low-glycogen workouts into their daily schedules.
A challenge for future investigations is to directly link
some of the acute exercise-induced molecular signaling
events in skeletal muscle that take place in response to
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1707
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
the greater metabolic loading imposed by various train-
ing interventions (i.e., altitude and low glycogen) to
dened performance-related outcomes that occur after
elite athletes undertake such practices.
2.6. A Time to Train, a Time to Compete?
Since the awarding of the Nobel Prize in Physiology or
Medicine in 2017 for the discovery that the molecular
clock is the primary mechanism underlying circadian
rhythms, there has been a dramatic increase in the num-
ber of scientic publications regarding circadian biology
and its impact on various aspects of human behavior,
including sporting performance. Circadian rhythms are
24-h (circa diem) oscillations in biological and meta-
bolic pathways. The circadian clock is cell autonomous
and present in most human tissues and organs and is
organized in a hierarchical manner, with the hypothala-
mic suprachiasmatic nucleus (SCN) functioning as the
master clockwith ne-tuningby clocks in peripheral
tissues (208211). Although light is the dominant zeit-
geber (time giver) for the SCN oscillator, which in turn
orchestrates rhythms in the peripheral organs/tissues at
appropriate phases, both the timing of exercise (212
216) along with the scheduling of meals (217220)can
impact circadian behavior (molecular underpinnings are
discussed in sect. 4.9).
Differences in the time of day for peak performance
for strength and anaerobic power as well as oxidative
capacity and endurance performance have been
reported in many, albeit not all, human studies (221
224). However, there are large interindividual differen-
ces in circadian rhythms, and the time of day for peak
performance is affected by many additional factors
including time since awakening, timing of precompeti-
tion meals, sleep quality, body temperature, hormone
levels, psychological habituation, motivation, and prior
muscle fatigue (225227). Accordingly, the effect of the
time of day of training on performance needs to be
placed in the context of an athletes chronotype. An indi-
viduals predisposition toward a preference for either
morning or evening can be classied into early chrono-
types (ECTs), late chronotypes (LCTs), or those in
between (intermediate chronotypes, ICTs) (228). ECTs,
sometimes referred to as larks,have signicantly ear-
lier sleep-wake cycles compared with LCTs (or night
owls), who function better later in the day. These differ-
ences are not only observed in sleep-wake cycles but
also multiple physiological (229), behavioral (228), and
genetic (230) oscillations that occur every 24-h period.
The implications for competition performance are not
entirely clear. Diurnal performance proles have been
studied between ECTs and LCTs to determine whether
there is signicant variation when individual aspects of
circadian timing are considered. These investigations
show clear differences in performance proles between
ECTs and LCTs, with LCTs exhibiting greater variation in
diurnal performance proles, particularly in the morning
(231). Interestingly, performance peaks can be shifted by
different measures such as active and passive warm-up,
caffeine, or training-competition time-of-day synchroni-
zation (225). Moreover, individual shifts in chronotypes
or time-of-day performance are observed (i.e., in older
athletes with a higher prevalence of morningnessin
training scheduling and work rates) (221).
The impact of exercise training at different times of the
day has been well studied in animal models and healthy
moderately trained humans, with the primary outcome
typically being a measure of exercise capacity, often a
laboratory-based task designed to mimic performance, or
a metabolic surrogate (232234). However, studies inves-
tigating the timing of exercise training in elite athletes and
the subsequent effect on performance outcomes are
scarce. Once again, we are left to generalize from inter-
ventions in healthy, almost exclusively male, non-elite
subjects until such gaps in the literature are lled. There
are several reports of greater increases in muscle mass
and strength after training late in the afternoon versus
early morning (221,233235), which is in line with the
generally higher peak forces attained in the afternoon/
early evening (236). Consistent with the enhanced reli-
ance on fatty acid oxidation in a fasted state in the early
morning in humans (237), there is a more robust meta-
bolicimpactofexerciseinthefastedstate(atthebegin-
ning of active phase in rodents) than in the fed state (at
the beginning of the rest phase) (214). Regardless, the
results are likely to have limited translational value for elite
athletes who typically undertake several workouts within
any 24-h period supported by round-the-clock eating pat-
terns necessary to meet the demands of training (81).
While elite athletes are informed of the venues, dates,
and times of major international competitions several
yearsinadvance,thenationsselectedtohostthe
Olympic Games and World Championships often adjust
competition times to accommodate and coincide with
prime-time viewing hours for North American television
audiences. At the recent Tokyo Summer Olympics, the
entire swimming program was ip-oppedsuch that
qualifying heats and seminals (normally held in the
morning) were scheduled for the evening and all nals
were swum in the morning. As circadian oscillations affect
physiological, psychological, and molecular mechanisms
resulting in varying physical performance capacities over
the day, both the timing and relative size of these effects
are important for optimizing sport performance at the elite
level. To determine the extent to which elite athletes are
affected by circadian uctuations in physical performance,
Lok et al. (238) assessed data from four Olympic Games
FURRER ET AL.
1708 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
(Athens 2004, Beijing 2008, London 2012, and Rio de
Janeiro 2016). The authors analyzed swimming perform-
ances, as these races are less likely to be inuenced
by confounding environmental effects (i.e., temperature,
humidity, wind speed) and have little reliance on equip-
ment that could induce variation within and between ath-
letes. Additionally, the water temperature in the pool is
required to be within a narrow range across Olympic ven-
ues, providing a cleansignalofdailyvariationinphysical
performance (238). Their analysis revealed that perform-
ance in Olympic swimmers was signicantly affected by
the time of day, with best performance occurring in the
late afternoon/early evening. The amplitude of the effects
of time of day was 0.37%, and in 40% of the nals this
effect was larger than the time difference between gold
or silver medal nishing times. Furthermore, time-of-day
effects exceeded the time difference between the silver
and bronze medals in 64% of the nals and the time differ-
ence between bronze and fourth place in 61% of the nals
(238). These data indicate that despite athletes incorpo-
rating both morning and evening workouts, endogenous
circadian clocks still exert a time-of-day effect on elite
swimming performance. Whether the application of circa-
dian or time-of-day principles can optimize training and
improve performance of these elite athletes remains to
be determined.
2.7. Training Strategies and Paradigms: Good,
Bad, or Indierent?
The identication of training strategies that consistently
enhance performance remains challenging because of
the multiple interdependent factors contributing to athletic
success. Consensus emerging from observational studies
reects the current practices in long (239), middle dis-
tance (146), or sprint (240) disciplines, but these are likely
to be modied with technological advances and
insights from coaches and science-driveninitiatives,
such as the sub-2 hourmarathon project. However,
training intervention studies are often limited to low partici-
pant numbers, with a reliance on a restricted pool of
young, male, college-educated, recreational or untrained
cohorts. Extrapolation to other demographics, including
women, underrepresented ethnicities, or elite athletes is
problematic. For example, signicant sex differences exist
in the response to both acute exercise and chronic training
adaptation (241243), and the effects of reproductive sta-
tus, endogenous and exogenous hormones, and the men-
strual cycle are underappreciated not only in research
studies but also in training program design and application
(244). Importantly, sex differences extend to many
other training-related factors, including muscle mass
and strength, injuries, and even training participation
rates (245). On occasion, understudied approaches
can lead to detrimental outcomes, as observed for the
transient hype surrounding the so-called benets of
cold-water immersion, whole body cryotherapy, and
other passive recovery strategies that in certain con-
texts can adversely affect recovery or performance
outcomes (246,247). New training strategies are often
based on observations of athletic performance in
extreme conditions, such as the high altitude of the
1968 Olympic Games held in Mexico City or the high
temperatures that were expected for the Tokyo 2020
Olympic Games (held in 2021 because of the coronavirus
pandemic). The former contributed to the widespread
study and adoption of high-altitude training, whereas the
latter was a primer to explore heat training as a modality
to improve performance not only in hot environments but
also in mild or cold temperatures. The potential mecha-
nisms for enhanced performance in thermoneutral envi-
ronments after heat exposure, improved thermotolerance,
enhanced heat dissipation, expanded plasma volume, ele-
vated hemoglobin mass, and other adaptations triggered
by heat exposure, are discussed in sect. 4.6.1 (248,249).
Safety is an obvious concern for such an intervention,
necessitating close monitoring of core body temperature
and cardiac-related parameters (250). In most cases,
many of the reservations about specic training strategies
stem from an inadequate understanding of the systemic,
organ/tissue, cellular, and molecular events that occur in
response to an acute exercise bout and how such infor-
mation translates into long-term adaptation. In sects. 3
and 4, we summarize the current knowledge of the physi-
ological, cellular, and molecular underpinnings of muscle
plasticity triggered by both endurance- and resistance-
based exercise.
3. PHYSIOLOGICAL AND CELLULAR
ADAPTATION TO EXERCISE TRAINING:
FUNCTIONAL RESPONSE
The cellular, tissue/organ, and whole body adaptations
that occur when exercise bouts are repeated over
months and years drive the phenotypic changes
observed in highly trained athletes. Such adaptations
include alterations in energy ux and metabolism, ber
type transformations, enhanced mitochondrial and capil-
lary density, and muscle hypertrophy, highlighting the
enormous plasticity of skeletal muscle (TABLE 1).
Although endurance training predominantly induces
numerous metabolic adaptations that match muscle
energy supply to demand and improve economy of
motion, athletes engaging in sports that require high peak
forces demonstrate marked changes in muscle morphology
and cross-sectional area (CSA). When training adaptation is
maximized in the face of favorable genetic predisposition,
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1709
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
extraordinary performances can be achieved, such as the
rst sub-2-h marathon by Eliud Kipchoge in 2019 (unofcial
record of 1:59:40.2), the 100 m time of 9.58 s in 2009 by
Usain Bolt, or the current world record holder Hafþ
or us
Bj
ornsson, who was able to deadlift 501 kg. These perform-
ances highlight the remarkable potential of skeletal muscle
to generate huge amounts of energy (adenosine triphos-
phate, ATP) for a sustained period in order to run at a speed
of 21.2 km/h for 2 h, to rapidly contract muscles to be able
to run at speeds exceeding 37 km/h for several seconds, or
to lift 500 kg. In this section, we discuss the adaptations
observed in elite athletes that allow such extraordinary
efforts.
3.1. Oxygen Transport and Maximal Oxygen
Uptake
Exercise of prolonged duration and/or high intensity
presents a major challenge to whole body homeostasis
and is associated with extensive perturbations in numer-
ous cells, tissues, and organs that are caused by, or are
a response to, the increased metabolic activity of con-
tracting skeletal muscles. To meet this challenge, multi-
ple integrated responses are rapidly engaged to blunt
the acute homeostatic threats generated by exercise-
induced increases in muscle substrate turnover and
oxygen demand (9). When repeated over time (i.e., exer-
cise training), there is a coordinated process of adapta-
tion that can be broadly categorized as either central
(nervous, respiratory, and cardiovascular systems) or
peripheral(skeletal muscle). However, such a simplis-
tic classication does not completely characterize the
interdependent nature of these processes. For example,
pulmonary oxygen diffusion, Hb levels, cardiac output,
vascularization of the muscle, as well as oxygen extrac-
tion and utilization by the muscle during oxidative phos-
phorylation (OXPHOS) all contribute to the endurance
training-induced increase in maximal oxygen uptake
(_
VO2max)(FIGURE 6)(291). _
VO2max therefore is a measure
of the combined capacities of the central nervous sys-
tem to recruit motor units, the pulmonary and cardiovas-
cular systems to deliver oxygen to contracting muscles
(including erythrocyte number and heme loading), along
with the ability of those muscles to extract and use oxy-
gen via oxidative metabolic pathways (9). At rest, whole
body oxygen consumption is 3.5 mL/kg/min, with
25% of this being taken up by skeletal muscle (251). In
untrained humans, _
VO2max is 10- to 15-fold greater
than resting values (i.e., 3050 mL/kg/min). In elite en-
durance-trained athletes, _
VO2max can be twofold higher
compared with non-athletes, with the highest _
VO2max
values being 96 and 80 mL/kg/min for male and female
endurance athletes, respectively (252,292,293). The
Table 1. Reference values of sedentary individuals and elite endurance athletes
Sedentary Elite Athlete References
Men WomenMen Women
_
VO2max, mL/min/kg <45 <40 7085 6075 (122,251265)
_
VEmax, L/min 120140 95 165185 125 (261,263,266273)
Stroke volume, mL/beat (251,253,265)
At rest 65 55 110 70
Maximum 100 70 150200 125
Cardiac output, L/min (251,253,265,274276)
At rest 563.54.5 563.54.5
Maximum 20 15 3040 25
Lactate threshold, % _
VO2max 60 60 7585 7585 (80,122,264,277)
Fiber type, % type I 4050 4050 >60 >60 (169,254,256,278284)
Capillary-to-ber ratio 1.52 Similar or slightly lower 2.53 insufcient data (169,254,255,278,285288)
Mitochondrial volume density, % 45 Similar or slightly lower 7.59 insufcient data (254,255,285,286,289,290)
Sedentary men and women are between 20 and 30 yr of age. All values of muscle tissue originate from vastus lateralis biopsies. _
VEmax, maximal exer-
cise-induced pulmonary ventilation; _
VO2max, maximal oxygen uptake. Less data are available for female athletes and sedentary control subjects.
FURRER ET AL.
1710 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
lower _
VO2max values for women are reected in the
female world records for endurance events, which are
typically 1012% slower (294). Such differences are mainly
due to the lower absolute and relative muscle mass in
women and the lower Hb levels (294). Differences in
_
VO2max are also observed based on the demands of the
sport and the associated mode of training. Elite male
cross-country skiers and rowers engage a larger propor-
tion of their muscle mass (upper and lower body) than
cyclists or distance runners during training and exhibit
higher _
VO2max values. The training-induced increases in
_
VO2max are largely conned to athletes competing in en-
durance-based events: power/strength-trained individuals
often have _
VO2max values similar to non-athletes (266).
Therefore, parameters affecting _
VO2max are differentially
impacted by the training prescription.
In most contexts, the limits of maximal oxygen con-
sumption are multifactorial and not attributable to any
single parameter (295). However, in general, exercise
performance is not limited by the respiratory system, as
its capacity exceeds demand during maximal exercise
(296). However, functional training-induced changes still
occur in athletes, including a greater strength and fa-
tigue resistance of the respiratory muscles resulting in
higher maximal voluntary ventilation (MVV) and forced
vital capacity (FVC) (266,296298). Maximal tidal vol-
ume is similar in athletes and non-athletes, with the
higher maximal exercise-induced ventilation brought
about by an elevated breathing frequency (266). Hb lev-
els are similar in highly trained endurance athletes and
untrained individuals (266) and are only elevated by alti-
tude training interventions, discussed in sect. 2.4. (299).
However, because of a training-induced increase in
blood volume, total Hb mass may be higher (300,301). In
addition to the increased strength of respiratory muscles,
endurance athletes have cardiac hypertrophy that is char-
acterized by greater left ventricular mass and relative wall
thickness resulting in a higher maximal and resting stroke
volume and a corresponding lower resting heart rate
(253,266). Although maximal stroke volume plateaus at
4050% of _
VO2max in untrained individuals, stroke vol-
ume increases until volitional exhaustion, contributing to
the augmented _
VO2max in elite endurance athletes (253).
In these individuals, maximum stroke volumes of 200 mL/
myonuclei ↑
MPS ↑
glycogen ↑
IMCL ↑
mitochondrial biogenesis ↑
capillary density↑
glucose↑
GLUT4 CD36
myoglobin↑
Whole body adaptation Skeletal muscle adaptation
fatty acids ↑
β-oxidation↑
OXPHOS ↑
Respiration
↑ strength & endurance of respiratory muscles
↑ breathing frequency
↑ maximal ventilation
Heart
cardiac hypertrophy
↑ resting/maximal stroke volume
↑ cardiac output
Blood
↑ Hb mass
↑ capillary density
Muscle
↑ slow oxidatve muscle fibers
↑ myoglobin
↑ substrate uptake and storage
↑ mitochondria
↑ OXPHOS
VO2max and endurance performance
Neural
↓ intracortical inhibitory interneurons
↓ antagonistic activity
Heart
cardiac hypertrophy
↑ resting stroke volume
Muscle
fiber hypertrophy
↑ myonuclei
Peak power
fiber hypertrophy
 

    



FIGURE 6. Whole body adaptations that contribute specically to higher peak power or endurance performance. Although mainly neural and muscular
adaptations improve peak power, for endurance performance various organs and tissues show major changes. To maximize _
VO2max and thereby endur-
ance performance, changes in respiratoryand cardiovascular function as well asadaptations in skeletal muscle are required. In skeletal muscle, the high
mitochondrial density, elevated substrate (i.e., fatty acids and glucose) uptake and storage, myoglobin content, and increased vascularization all contrib-
ute to the elevated performance of endurance athletes. Strength training-induced adaptations include increased muscle protein synthesis (MPS) result-
ing in ber hypertrophy and optimally myonuclear accretion. CD36, platelet glycoprotein 4; GLUT4, glucose transporter type 4; IMCL, intramyocellular
lipids. See GLOSSARY for other abbreviations. Sport icon vectors were created by ibrandify/Freepik. Image created with BioRender.com, with permission.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1711
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
beat have been reported, indicating that a cardiac output
of up to 3540 L/min could be reached, a gure almost
double that observed in non-athletes (251,274). Whereas
left ventricular mass and resting stroke volume are similar
in power and endurance athletes, _
VO2max is not elevated
in power athletes, likely because of the increased maximal
stroke volume and oxygen pulse per stroke volume after
endurance but not resistance training (266). Finally, vascu-
larization of skeletal muscle also contributes to _
VO2max
(154). Capillary density in all ber types is 50% higher in
elite endurance-trained compared with nontrained individ-
uals (254,278), and athletes with superior muscle vascula-
rization are even more fatigue resistant compared to
athletes with similar _
VO2max values (80). The endurance
training-induced vascularization occurs rapidly, with 6 wk
of intense training being sufcient to elevate capillary den-
sity and capillary-to-ber ratio (molecular mechanisms driv-
ing this adaptation are described in sect. 4.5.3) (255,302).
In contrast, capillary density in power/strength-trained ath-
letes does not increase with training and may be lower
than that in untrained individuals because of the ber hy-
pertrophy (278,303). Collectively, central adaptations
occur at multiple levels and play an important role in the
high _
VO2max in elite athletes.
3.2. Neuromuscular Control and Force Generation
In contrast to the high _
VO2max required for optimal en-
durance performance, many sports require high power
generation, including sprint events (running, swimming,
cycling, and rowing) and weightlifting, powerlifting, and
throwing events. Maximal performance in elite strength/
power athletes is 1520% lower in females than in
males (304), because of differences in lean body mass
between men and women (168,305). In line with the
lower lean mass, fat mass of female athletes is about
twofold higher compared to men with similar body mass
(304). With increasing age, differences in maximal per-
formance in terms of world records in female and male
masters athletes become greater, and records are
3050% lower in women, mirroring the sex differences
observed in untrained and recreationally trained individ-
uals (304).
Voluntary muscle contraction is a complex task requir-
ing a highly coordinated interplay on multiple levels
including supraspinal structures, spinal cells, afferent
feedback and efferent input, and the motor unit (306).
The motor unit consists of the soma, dendrites, and
axon of the motor neuron as well as the innervated mus-
cle bers. The force-generating capacity of the muscle
is determined by the number of activated motor units,
the discharge rate (also described as ring frequency
or rate coding) of the motor neuron, and the size and
contractile properties of the activated muscle bers. To
initiate muscle contractions, the central nervous system
sends commands to the motor neurons located in the
ventral horn of the spinal cord (307). Motor neurons inte-
grate the signal from a number of different regions and
nuclei in the cortex and brain stem, interneuron circuitries,
as well as the peripheral sensory input from afferent bers
located in the muscle spindles and Golgi tendon organ
into an action potential. The action potential propagates
along the axon of the motor neuron to the innervated
muscle bers and, through acetylcholine receptor activa-
tion and excitation-contraction coupling (ECC), results in
mechanical output by the muscle. In comparison to central
nervous system synapses, the neuromuscular junction
(NMJ) has a very high safety factor, and a nerve action
potential results in an end-plate potential (EPP) of a local
depolarization of 3040 mV, which is higher than
required to elicit an action potential in the muscle. Several
morphological and functional parameters also contribute
to this high safety factor (308310): rst, an extraordinarily
large size of the synapse, 100- to 200-fold bigger com-
pared with central nervous synapses in the mouse, and
thus ample interaction surface; second, a high density
of voltage-gated Ca
21
channels in the active zones,
coupled to a high concentration of acetylcholine in a syn-
aptic vesicle; third, the number and density of acetylcho-
line receptors, and the concentrated localization at the
crest of postsynaptic folds, adjacent to voltage-gated Na
1
channels (Nav1.4) in the corresponding troughs; and fourth,
the strong enzymatic activity of acetylcholinesterase in the
synaptic cleft for rapid removal of acetylcholine and
thereby prevention of repeated activation of individual ac-
etylcholine receptor channels in response to a single
action potential in the motor neuron. Together, these prop-
erties lead to an all-or-noneactivation (as rst described
by Henry Pickering Bowditch in 1871 for cardiac muscle,
later expanded to skeletal muscle), meaning that once the
stimulus threshold for an action potential in the motor neu-
ron is reached (based on the integration of different incom-
ing signals), an action potential and contraction in the
muscle ber is inevitably triggered. The frequency of acti-
vation (rate coding) of the muscle berisimportantforthe
generation of force (311). In a single muscle twitch, Ca
21
reenters the sarcoplasmic reticulum and ber relaxation
becomes complete. A more frequent stimulation results in
wave summation, and thus greater force, ultimately maxi-
mizing in a tetanus, in which twitches overlap and no
relaxationcanoccur.Inrodents,exercise-induced
NMJ remodeling has been observed affecting mor-
phology and function of this synapse (312,313). For
example, endurance training boosts the amount of
neurotransmitter released per action potential, con-
comitantly with an upregulation of acetylcholinester-
ase (313,314). Furthermore, an enlargement in the
interaction surface is achieved by a modulation of the
FURRER ET AL.
1712 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
number and length of nerve terminal branches, coupled to
an elevation in the total area occupied by presynaptic neu-
rotransmitter vesicles and postsynaptic acetylcholine recep-
tors (314). Similar adaptations are observed in genetic
mouse models of endurance training, with corresponding
changes in neuromuscular transmission properties (315).
Notably, however, size, complexity, and fragmentation of
murine and human NMJs can differ substantially (316). Data
describing training-induced NMJ plasticity can therefore
only be extrapolated to humans with caution, in particular
since corresponding interrogations in humans are lacking.
Whereas hand muscles have fewer motor units than
large limb muscles, the number of motor units of different
limbmusclesvariesandisnotalwaysrelatedtomuscle
size (317). In contrast, the average innervation number
(number of muscle bers innervated by a single motor neu-
ron, also called motor unit size) strongly correlates with
muscle size (317). Innervation numbers, even within one
muscle, can range from tens to thousands, and thereby
enable diverse actions such as ne-tuning of the move-
ment or high force generation, respectively (317,318).
Whereas slow type I muscle bers (discussed in sect. 3.3)
are mostly part of motor units with a low innervation num-
ber, motor units with a high innervation number often con-
nect to fast type II bers (317). The motor neurons of these
different motor units exhibit considerable morphological
and functional differences (FIGURE 7). For example, motor
neurons innervating type II muscle bers in general have
larger somas and more dendrites as well as a larger axonal
diameter sizes, the latter enabling faster conductance ve-
locity (306,319). The physical dimensions of the motor
neuron somas contribute to the determination of the
recruitment threshold (320). Thus, the larger surface area
and high number of ion channels in fast motor neurons
result in a lower input resistance compared with the small
surface area with fewer ion channels of slow motor neu-
rons (320). According to Ohmslaw(V=IR), the same
synaptic input thus induces greater changes in the mem-
brane potential of small motor neurons (with a higher re-
sistance) compared with large motor neurons (with a lower
resistance). Consequently, small motor neurons reach the
ring threshold with less synaptic input compared with
theirlargercounterparts.Thisorderlyrecruitmentwas
showninanimalpreparationsbyHenneman(320), and
according to Hennemans size principle small motor neu-
rons innervating slow type I muscle bers are recruited
rst, subsequently followed by larger motor neurons inner-
vating type IIA and nally IIX bers (306). This leads to a
gradual and smooth increase in muscle force (graduation
of contraction) and a predominant activation of slow and
?
antagonists
motor unit recruitment
?
?
inhibition of
afferent feedback
and/or
spinal interneurons
intracortical
inhibition
corticospinal
excitability
excitatory
drive low threshold
motor unit
high threshold
motor unit
NMJ
remodeling
fiber size
firing frequency
agonists
synergists
stabilizers
FIGURE 7. Neuromuscular adaptation totraining. The number of activated motor units, their ring frequency, as well assize and contractile properties
of the muscle bers determine total force-generating capacity. In trained individuals, neural adaptations include an increased excitatory drive that can
lead to an elevated ring frequency and higher number of activated motor units. In addition, the enhanced activation of agonists, synergists, and stabil-
izers together with the reduced coactivation of antagonists contribute to the increased force production after training. For many of these adaptations
only data in rodent models exist, and/or controversial ndings in humans have been reported. NMJ, neuromuscular junction. Illustration of person was
created by kjpargeter/Freepik. Image created with BioRender.com, with permission.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1713
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
fatigue-resistant small motor units until these are over-
whelmed by strong, powerful movements necessitating
the recruitment of fast-twitch, high peak force-generating
bers (306). Although the size principle holds true for spe-
cic laboratory conditions, physiological systems that
include excitatory and inhibitory inputs are much more
complex, and it is debatable whether motor units are
always recruited in a graded manner (319,321). Recent evi-
dence suggests that during slow, ramping movement
orderly recruitment occurs, whereas high-frequency, sinu-
soid types of contractile activity are not following the size
principle (322). Moreover, motor unit recruitment is
affected by the length of the muscle (322). This implies
that, during ballistic training, recruitment is observed in a
more selective manner corresponding to the functional
movement according to the neuromechanical matching
principle (319,321). However, even during slow ramping
movement, as might be encountered in weightlifting, fast
motor units can be recruited with submaximal load (138,
323,324). In fact, a number of studies have demonstrated
that even low-load exercises such as 30% 1RM result in
recruitmentoftypeIandIIber if completed until failure,
inducing a hypertrophic response similar to high-load train-
ing (138,323325). This nding might be explained by the
observation that with increasing muscle fatigue more
motor units are being recruited to meet the same force
output,evenwhenlowerloadsareused(138,323,324).
Therefore, the recruitment of slow and fast motor units is
dependent not only on the applied load and generated
force but also on the fatigue state, contraction velocity,
and length, as well as potentially other parameters that are
extracted from skeletal muscle in different exercise proto-
cols and paradigms. Some of these factors, such as velocity
and length, might induce selective rather than the orderly
recruitment of motor units postulated in Hennemanssize
principle. In summary, motor unit engagement, recruitment,
and fatigue are still poorly understood. Moreover, the plas-
ticity of this structure in training and the specicadaptations
in elite athletes are largely unknown. Of note, based on the
transcriptional prole, even more distinct motor neuron
pools exist than the classically dened three types, fast fati-
gable, fast fatigue-resistant, and slow (326). The functional
relevance of this more ne-grained specication remains
unknown.
3.2.1. Neural adaptations.
Increased force generation can be achieved by neural
adaptation, muscle hypertrophy, and/or intrinsic changes
in contractile properties (strength/power per unit of muscle
mass). MVC is substantially elevated in resistance training
beforeanyincreaseinmuscleCSA,suggestingthatneural
adaptations mainly contribute to the strength gains in the
initial phase, followed by structural changes within the
muscle. The activity of intracortical inhibitory interneurons
is lower in trained individuals, whereas data on corticospi-
nal excitability are equivocal and likely unaffected by train-
ing (327329). The reduction in intracortical inhibition in
response to strength training induces a higher excitatory
drive, which may contribute to elevated strength in these
individuals. With augmenting force generation, the number
of active motor units as well as their activity in terms of dis-
charge rate increase (318,330). Accordingly, the higher de-
scending neural input in strength-trained individuals could
explain the elevated discharge rate and adequate activa-
tion of motor units, culminating in the observed enhanced
voluntary activation and maximal force (318,331,332). In
addition, the elevated neural drive might be important for
explosive power, as a fast recruitment and a high discharge
rate of motor units are important for the rate of force devel-
opment (330). The lower recruitment threshold observed
after strength training suggests that, in addition to the
changes in neural drive, intrinsic properties of motor neu-
rons may be altered by training (331). The timing of the
action potentials discharged by concurrently active
motor units of strength athletes appears to exhibit
greater synchronization than that in untrained individ-
uals, even though it is debatable whether these adap-
tations in intramuscular coordination contribute to
strength gains (333). Intermuscular coordination is
also improved at several levels. Besides the recruit-
ment of agonists, stabilizers/xators, and neutralizers/
synergists, reducing the coactivation of antagonists
signicantly contributes to the maximal voluntary acti-
vation and force generation in highly trained athletes
(332334). Whether these training adaptations
evoked by intermuscular coordination are primarily
mediated by disinhibition of supraspinal signals,
altered activity of Renshaw cells and other spinal inter-
neurons, and/or adaptations in afferent feedback,
such as decreased stretch inhibition by the proprio-
ceptive system, is unclear. Nevertheless, the impor-
tance of early neural and neuromuscular adaptations
and the concomitant optimization of intra- and inter-
muscular coordination before structural changes of
themuscleinresistancetrainingadaptationisirrefuta-
ble (FIGURE 7)(329,335,336). The corresponding
neural changes in endurance training are less well
characterized but could contribute to improved run-
ning economy, decreased fatigability, and other inter-
related parameters (337339).
3.2.2. Muscle hypertrophy.
Powerlifting or hammer throw athletes rely on the gener-
ation of instantaneous maximal peak forces, whereas
sports that involve short-duration sprints require high
contractile velocity of muscles over several seconds.
FURRER ET AL.
1714 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
Nonetheless, an increase in muscle mass to generate
high contractile forces is a common goal for these ath-
letes. Indeed, the muscle volume of the lower limbs of
both sprinters and other strength-trained athletes is
higher compared with endurance athletes and untrained
individuals (340,341). The training-induced gain in mus-
cle CSA is not evenly distributed along the length of the
muscle ber but occurs predominantly in the midbelly
region of the muscle, which explains why the percent
increase in CSA can exceed that of muscle volume (335,
342). In humans most muscles are pennate, in which the
increase in physiological CSA of the muscle, reecting
the radial growth of the myober, can diverge from the
increase in anatomical CSA. To optimize the limited
space of the aponeurosis, a larger CSA is usually accom-
panied by a steeper pennation angle, which is greater
in highly trained strength athletes compared with
untrained individuals (343) and strongly correlates with
muscle thickness (344). In contrast, in elite sprint athletes
(100 m sprinters and sprint cyclists), muscle thickness is
increased without changes in the pennate angle (345,
346). As these observations in elite athletes are all cross
sectional, it remains to be determined whether the archi-
tectural differences contributing to peak performance
are the result of long-term training adaptation or genetic
predisposition. Despite these ndings, the contribution
of the pennation angle and other architectural proper-
ties to muscle functionality and power generation
remains contentious (347,348).
Of note, hypertrophy of myobrillar and sarcoplasmic
compartments has been described, and the relative
impact on muscle mass and strength gains remains equiv-
ocal, similar to the importance of conventionalhypertro-
phy with a proportional increase in myobrillar protein
content and tissue growth compared with unconven-
tionalhypertrophy, for example achieved by myobrillar
packing preceding an increase in ber size (349).
Nevertheless, in most cases resistance training induces
radial growth of the muscle, resulting in a higher CSA
(mechanisms underlying this response are described in
sects. 4.1 and 4.34.5) (342,350). The expansion of myo-
brillar protein resulting in ber hypertrophy might contrib-
ute the most to enhanced force-generating capacity of a
muscle ber. Within the muscle ber, 80% of the volume
consists of myobrils that are composed of sarcomeres,
the contractile units of the myobril (351). The thin actin
and thick myosin laments constitute the two major active
components of the sarcomeres responsible for muscle
contraction. Upon Ca
21
binding to troponin C (TnC), tro-
pomyosin conformation changes to expose the myosin-
binding site on the actin lament. The thick myosin la-
ments, the force-generating elements of the sarcomere,
bind to actin and induce the sliding of actin laments
along the myosin, resulting in muscle shortening. The
addition of sarcomeres in parallel rather than in series
causes an increase in ber diameter (342). In line with the
high potential of type IIA bers to increase CSA and force
generation (352), hypertrophy predominantly occurs in
type IIA bers in elite strength-trained athletes (303,350).
Besides radial growth, inclusion of additional sarcomeres
in series leading to increased fascicle strength has been
reported (342). Limited data are available regarding the
longitudinal growth of the muscle in response to resist-
ance-based training, although there is evidence that fasci-
cle length can increase (335,343). For example, a longer
fascicle length is observed in elite sprinters (345,346),
which is positively correlated with performance times
(343,353,354). The longer fascicles could contribute to a
greater shortening velocity of a pennate muscle and
thereby enhance sprint performance (353). However, the
totalnumberofsarcomeresinseriesinamuscleber and
the effects of training are difcult to determine in humans,
and the few studies in rodents revealed mixed results
(342). Moreover, recent evidence indicating a mesh-type
network of branching sarcomeric structures instead of
individual sarcomeres existing in separated tubes further
complicates the interpretation of changes in sarcomere
numbers in series and in parallel (355).
Despite the fundamental contribution of radial muscle
growth to maximal power output, the relationship between
force generation and muscle CSA is not linear, emphasiz-
ing the contribution of other factors (356). One possibility
could be that muscle quality rather than size is enhanced,
resulting in a higher specic force (force per CSA) (352).
For example, the specic force of type I bers has been
shown to increase in response to resistance exercise
(352). Additionally, changes in ber type distribution could
enhance specic force capacity, since both the force-gen-
erating capacity per myosin head as well as the fraction of
attached myosin heads are higher for fast myosin heavy
chain isoforms (357). Despite these changes, the increase
in overall muscle strength is superior compared with the
integration of single-ber strength gains, indicating that
optimalstrengthgainsoccurwhenbothneuralaswellas
muscular adaptations take place (352).
The increase in myobrillar proteins is often, but not
always, accompanied by elevation of the number of myo-
nuclei, potentially to optimize the hypertrophic response.
The syncytial nature of muscle cells has been hypothe-
sized to be due to the limited capacity of (myo)nuclei to
provide transcripts for a certain volume of the cytoplasm,
dened as the myonuclear domain (358). According to
this hypothesis, the upper limit of the myonuclear domain
is determined by the maximal transcriptional capacity of
myonuclei. Once this ceiling is reached, the number of
myonuclei was thought to be increased by the fusion of
satellite cells to muscle bers to maintain a relatively con-
stant DNA-to-cytoplasm ratio (reviewed in Ref. 358), a
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1715
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
concept that has subsequently been rened and cor-
rected (359). Moreover, in humans hypertrophy has also
been reported in the absence of myonuclear accretion
(360,361). Furthermore, although the addition of myonu-
clei is greater when hypertrophy exceeds 22% of size
gain, it also occurs when hypertrophy is <10% (362).
According to a recent meta-analysis, a denitive hypertro-
phy thresholdrequired for myonuclear accretion remains
tenuous, and thus the threshold hypothesis appears prob-
lematic (362). Nevertheless, cross-sectional data from ath-
letes often show hypertrophy to be positively correlated
with myonuclear number (363). For example, in elite
powerlifters, both the size of the muscle bers as well as
thenumberofmyonucleiarehighercomparedwithnon-
athlete control subjects, and the gradient of the correlation
curve suggests that the myonuclear domains are higher in
largetypeIImusclebers of powerlifters compared with
those of untrained individuals (364). The use of anabolic
steroids and testosterone in powerlifters results in a dispro-
portional increase in muscle size and number of myonu-
clei, thereby leading to a larger myonuclear domain (365).
Although these data indicate a certain degree of myonu-
clear domain plasticity that could be explained by the
reserve capacity of the myonuclei to boost transcriptional
output (366), in most cases the myonuclear domains are
still within the reported range, also highlighting individual
heterogeneity.
In summary, strength-trained athletes have a pro-
nounced increase in muscle volume largely underpinned
by the specic hypertrophy of type IIA bers and accom-
panied by an increase in pennate angle, resulting in sub-
stantial gains in force-generating capacity. In addition to
the elevated muscle volume, the fascicles of sprint ath-
letes are elongated, contributing to higher shortening
velocity of the muscle and greater power generation.
Although neuromuscular adaptations together with mus-
cle hypertrophy explain the enhanced muscle perform-
ance in highly trained athletes, these adaptations differ
among athletes in a discipline-dependent manner. For
example, the musculature of strength-trained athletes has
a larger CSA and can produce greater maximal forces,
whereas power-trained athletes have slightly lower peak
forces but display a faster rate of force development (341).
These differences are most pronounced within the rst 50
ms of a contraction (341,367). The enhanced explosive
(i.e., high shortening velocity) and maximal forces in ath-
letes result in remarkable power output. In male power
athletes, mean peak power, as assessed by a counter-
movement jump, ranges between 50 and 65 W/kg, with
maximal values up to 85 W/kg for men and 70 W/kg for
women (252,303,368,369). The acceleration power
attained at the start of a 100 m sprint was estimated to be
2,392 W or 30.3 W/kg for men and 1,494 W or 24.5 W/kg
for women (252). It was calculated that Usain Bolt
reached a power output of 2,750 W (29.3 W/kg) during his
100mworldrecordin2009(252). In contrast to these re-
markable power metrics for strength/sprint-trained athletes,
power output (as measured by jump performance) remains
largely unchanged in endurance-trained athletes (266). It
is clear that the adaptive response to training stimuli is
event-specic in terms of muscle hypertrophy and neuro-
muscular adaptation, resulting in distinct performance
characteristics between different sporting disciplines.
3.3. Fiber Type Distribution
Distinct properties of the muscle further contribute to
the performance of elite athletes, such as ber type dis-
tribution. In addition to distinct innervation and recruit-
ment (as described in sect. 3.2), intrinsic properties of
muscle ber types diverge in a multifaceted manner
(155,357,370372). Muscle ber types can be classied
according to their predominantly expressed isoform of
the myosin heavy chains, which are the molecular
motors of the myobrils and vary in the relative actin-
activated ATPase activity that correlates with contraction
velocity (357). Differences in ber type distribution are
accordingly observed in muscles of endurance- or
strength-trained athletes and considerably contribute to
sports-specic performance. The three myosin heavy
chain isoforms expressed in human muscle, type I
(encoded by the MYH7 gene located on chromosome
14, 14q11.2), type IIA (encoded by MYH2, on chromosome
17, 17p13.1), and type IIX (encoded by MYH1,adjacentto
MYH2 on chromosome 17, 17p13.1), have distinct me-
chanical properties conferring differences in contractile
velocity and force production. Type IIX bers generate
the highest force and have the fastest shortening veloc-
ity, resulting in high peak power, and are classied as
fast-twitchbers (357,373). The enhanced force-gener-
ating capacity is attributable not only to the larger size of
type II bers but also to intrinsic differences (i.e., higher
force-generating capacity of fast myosin heavy chain iso-
forms and a larger fraction of attached myosin heads),
resulting in a higher specic force of type II bers, which
is observed in untrained individuals as well as elite ath-
letes (357,374376). In addition to greater power-gener-
ating capacity, fast type II muscle bers exhibit shorter
half-relaxation time due to differences in Ca
21
transient
kinetics (357). In response to an action potential, Ca
21
release in fast murine muscle bers is threefold higher
compared with slow bers, likely because of the greater
abundance of the Ca
21
release channel ryanodine recep-
tor 1 (RYR1) (357). The inhibitory effect of intracellular
Mg
21
concentrations on Ca
21
release is lower in slow
muscle bers, possibly contributing to the higher fatigue
resistance of these bers, as Mg
21
levels rise during fa-
tigue (357). Fiber type-specic differences in Ca
21
FURRER ET AL.
1716 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
transient related to the faster decline in cytoplasmic con-
centrations are determined by the sequestration of Ca
21
tobindingandbufferproteinssuchastroponinC(TnC),
parvalbumin, and calmodulin, as well as the reuptake of
Ca
21
into the sarcoplasmic reticulum (SR) by the sarco-
plasmic/endoplasmic reticulum Ca
21
-ATPase (SERCA)
pumps (357). As the TnC isoform expressed in fast muscle
bershasfourCa
21
-binding sites compared with the
three in the isoform expressed in slow bers, Ca
21
bind-
ingisenhancedinfastmusclebers. Moreover, the faster
uptake of Ca
21
by the SR is determined by the increased
SR volume and surface area as well as the higher density
of SERCA pumps in fast bers. Of note, the SERCA iso-
form expressed in fast bers (SERCA1 vs. SERCA2 in slow
bers) is more sensitive to changes in ADP concentration,
which rises during muscle fatigue. As a result, Ca
21
-pump
and -leak rates of SERCA1 are more affected by metabolic
stress (i.e., more reduced and more increased, respec-
tively) compared with SERCA2 in slow bers (357). In addi-
tion, calsequestrin (CASQ) that binds Ca
21
within the SR is
found in greater abundance in fast compared with slow
bers, thereby providing an increased capacity to bind
free Ca
21
. Taken together, differences between Ca
21
transient and cross-bridge kinetics in slow and fast muscle
bers contribute to the distinct contractile properties
(FIGURE 8). Accordingly, it is not surprising that the
energy demand of these ber types is different during
maximal isometric contraction. At rest, energy expenditure
in muscle is relatively low, 0.008 mM/s of ATP turnover,
and mainly used for the Na
1
-K
1
-ATPase in the sarco-
lemma as well as protein synthesis (357). However, as
muscles start to contract, the energy demand is substan-
tially elevated by the myosin ATPases of the molecular
motor for cross-bridge cycling (70% of overall ATP
glycolytic
muscle fiber
SR
SERCA1
MyHC-IIX
RyR1 CASQ
β-oxidation
IMCL
myoglobin
glycogen
SR
MyHC-I
SERCA2
RyR1
CASQ
Time
Force
Time
Force
oxidative
muscle fiber
Contractile characteristics
Ca2+
Ca2+-binding
proteins
ATP
glycolytic
oxidative
Ca2+
Ca2+-binding
proteins
ATP
glycolysis
pyruvatelactate
OXPHOS
Fatigue glycolysis
pyruvate
β-oxidation
OXPHOS
OXPHOS
PPPPPP
PPP
PPP
PPP
FIGURE 8. Contractile and metabolic properties of a strength/power-trained and an endurance-trained athlete. Characteristics of fast and powerful
muscles of strength/power athletes that are accompanied by a more fatigable muscle as compared with endurance-trained muscles with elevated oxi-
dative capacity that are more fatigue resistant. ATP, adenosine triphosphate; CASQ, calsequestrin; IMCL, intramyocellular lipids; MyHC, myosin heavy
chain; OXPHOS, oxidative phosphorylation; RyR1, ryanodine receptor 1; SERCA, sarcoplasmic/endoplasmic reticulum Ca
21
-ATPase; SR, sarcoplasmic
reticulum. Illustrations of people were created by kjpargeter/Freepik. Image created with BioRender.com, with permission.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1717
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
consumption) and by the SERCA pumps for Ca
21
reup-
take (30% of ATP consumption). Hence, during maximal
isometric contractions ATP turnover rates in human type I,
IIA, and IIX bers can increase up to 1,000-fold to 1.7,
4.7, and 7.2 mM/s, respectively (357). These ber-specic
differences underscore the remarkable capacity of fast
muscle bers to turn over ATP and enable the generation
of high peak power.
To meet the distinct energy demands of different
sport activities, slow and fast muscle bers exhibit diver-
gent metabolic properties. Type I bers are highly oxida-
tive, whereas type IIX bers are more glycolytic (357),
with the fast type IIA bers of an intermediate phenotype
containing high amounts of oxidative as well as glyco-
lytic enzymes. In line with the considerable energy
demand of fast glycolytic bers, glycogenolytic activity is
elevated in these bers, and during short maximal con-
tractions ATP generation via glycolysis is double that of
slow muscle bers (357), with a corresponding rapid rate
of fatigue. In contrast, rates of fatty acid oxidation via the
b-oxidation pathway are two- to threefold higher in slow
bers, with the lower ATP demand of these bers met
by mitochondrial respiration over prolonged periods
(357). The increased oxidative capacity of type I bers is
determined by the higher mitochondrial volume, account-
ing for 6% of the bervolumeintypeIcomparedwith
4.5% and 2.3% in type IIA and IIX bers, respectively, as
well as the greater density of the mitochondrial cristae
and enzymatic activity (357). Additionally, the elevated ox-
ygen demand in slow bers is met by a higher capillary
density and 50% increased myoglobin content compared
with fast bers (154,357). Along with the capillary-to-ber
ratio,thepercentageoftypeIbers thus is a strong pre-
dictor of endurance capacity (377). The capillary density
not only provides greater oxygen delivery and energy
substrates but also promotes rapid removal of by-prod-
ucts of sustained contractile activity (i.e., ammonia or lac-
tate). Triglyceride stores are substantially higher in type I
bers compared with type II bers (0.5% of the ber vol-
umeintypeIbers compared with <0.1% in type IIX
bers), whereas conicting results have been reported
regarding glycogen stores (279,357,378). Although ber
type-specic differences in glycogen content are not
always observed between fast and slow muscle bers
(379,380), several studies report that glycogen concentra-
tion in type II bers is 1530% higher compared with type I
(357,381385). It is unclear whether this is due to differen-
ces in methodology or whether the training status has an
impact on these ndings, since the glycogen content is
similar in type I and type II bers in elite athletes (386).
Collectively, the ber type-specic differences in con-
tractile and metabolic properties reected by the inverse
relationship between force generation and oxidative
capacity of type I, IIA, and IIX bers result in low force-
generating, fatigue-resisting bers and high force-gen-
erating, fatigable bers, respectively. These properties
emphasize the important contribution of ber type distri-
bution to endurance- or strength/power-based activities.
In untrained males, the vastus lateralis comprises 40
50% type I bers (254,256,278,387), though the rela-
tive ber type distribution depends on the site of biopsy
(388). Elite endurance athletes typically present with
>60% type I bers, with extremes of >90% (169,254,
256,279,280,389). These cross-sectional data, how-
ever, fail to provide evidence of whether prolonged en-
durance training induces a shift in ber distribution, or if
elite endurance athletes are successful because of innate
ber type patterns.
In contrast to endurance athletes, well-trained strength
and power athletes tend to have a ber type distribution
that resembles that of untrained individuals, at least in
terms of overall glycolytic (type II) compared with oxida-
tive (type I) bers, with some extremes toward a lower
abundanceoftypeIbers (278,303,368,390). A signi-
cant shift from IIX to IIA bers has been reported in
strength- and power-trained athletes, with the absolute
area of fast muscle bers substantially larger because of
specic hypertrophy of type IIA bers (278,303). In other
studies, a preservation of type IIX bers in response to
sprint or plyometric training has been found, together
with a shift from type I to type IIA (372). Accordingly,
sprinters seem to have a lower proportion of slow type I
bers (281,391). As for elite endurance athletes, it is
unclear whether differences in ber type distribution in
power athletes and sprinters are due to preexisting ber
type patterns or whether the training-induced shift is
affected by different training paradigms. It also remains
perplexing how a ber type shift is brought about in terms
of the temporal and spatial coordination: is there a simul-
taneous shift in metabolic and contractile properties com-
bined with changes in the motor neuron phenotype, or
does one follow the other? Collectively, it is evident that
the unique characteristics of type I, IIA, and IIX bers
regarding oxidative and force-generating capacities are
instrumental for sport-specic demand and thereby con-
tribute to the achieved peak performances in athletes.
3.4. Energy Metabolism and Oxidative Capacity
In skeletal muscle, the three main sites for ATP consump-
tion are the Na
1
-K
1
-ATPases of the sarcolemma for mem-
brane excitability, the SERCA pumps of the SR membrane
for Ca
21
reuptake, and the myosin ATPases for cross-
bridge cycling. Because of the increased elevation in ATP
utilization by myosin ATPases and SERCA pumps associ-
ated with intense contractile activity, the high ATP
demand is a major bioenergetic challenge to the con-
tracting myobers (357). Given that intramuscular ATP
FURRER ET AL.
1718 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
stores are relatively small, metabolic pathways re-
sponsible for ATP restoration must be rapidly acti-
vated such that ATP levels closely match demand.
During sprint events lasting <20s,thecreatinephos-
phate system and anaerobic glycolysis are the main path-
ways engaged in ATP resynthesis (392). Although the
contribution to ATP provision during a 10-s sprint is similar
between the creatine phosphate system and anaerobic
glycolysis (53% and 44%, respectively), this is substantially
changedduringa30-ssprint(23%and49%,respectively,
with 28% from mitochondrial respiration) at a time when
phosphocreatine stores are mostly depleted (392,393).
In contrast, to meet the elevated and sustained energy
demand for prolonged exercise (several minutes to sev-
eral hours), large amounts of ATP are generated by aero-
bic mitochondrial respiration, with the metabolism of
glucose, fatty acids, ketone bodies, and lactate resulting
in a higher energy expenditure (392). In extreme endur-
ance events such as the Ironman triathlon (3.8 km swim-
ming, 180 km cycling, 42.2 km running), rates of energy
expenditure 8- to 10-fold above resting metabolic rate
(RMR) can be sustained for 1015 h (394). To enable such
performances, adequate oxygen uptake, delivery, and
extraction concomitant with skeletal muscle ATP genera-
tion using large amounts of fatty acids must be tightly
coordinated.
In addition to the higher _
VO2max values observed in
elite endurance-trained athletes, these individuals also
have greater efciency or economy of motion (i.e., a
lower oxygen cost at any given work rate or speed of
movement) combined with a greater fractional utilization
of _
VO2max (i.e., the ability to use a greater proportion of
their higher _
VO2max). Collectively, these attributes are
associated with increased rates of fatty acid oxidation, a
slower rate of depletion of muscle glycogen stores, and
enhanced lactate kinetics (395,396). For example, maxi-
mal rates of whole body fatty acid oxidation in elite ath-
letes are twofold higher at submaximal workloads
compared with untrained individuals (0.6 g/min vs. 0.3 g/
min) (254) and can be reached at much higher intensities
(both relative and absolute) compared with untrained indi-
viduals (4560% vs. 35% of _
VO2max). A retoolingof
themusclebyadaptationtoahigh-fatdietwhileunder-
taking vigorous training can more than double these rates
in world-class endurance-trained athletes to values
exceeding 1.5 g/min at work rates of 70% of _
VO2max
(397). During exercise exceeding 80% of _
VO2max,the
availability or rate of appearance of fatty acids in the cir-
culation limits the oxidation of fat-based fuels by skeletal
muscle (398). In line with a superior ability to oxidize fat-
and spare carbohydrate-based fuels during exercise, the
onset of blood lactate accumulation, xed at a concentra-
tion of 4 mM (sometimes arbitrarily dened as lactate
threshold)(255), is observed at both higher relative and
absolute exercise intensity in endurance-trained athletes
compared with untrained individuals (80,257). Therefore,
in addition to the higher _
VO2max and improved capillary
density, training-induced adaptations in skeletal muscle
(i.e., the ability to oxidize fat-based fuels and spare muscle
glycogen along with a concomitant reduction in lactate
appearance) enable elite endurance athletes to sustain
high absolute work rates or speeds and resist fatigue
for prolonged periods (80,267,395). These adaptive
changes are discussed in sects. 3.4.1 and 3.4.2.
3.4.1. Mitochondrial adaptations.
During prolonged submaximal exercise with adequate
oxygen availability, ATP is synthesized by OXPHOS in the
electron transport chain located in the cristae of the inner
mitochondrial membrane. To optimize mitochondrial bio-
energetics and meet the increased energy requirements
during prolonged exercise, highly coordinated adaptive
processes enhance the quality as well as the quantity of
mitochondria. For example, mitochondrial number and
morphology are altered by replication of mitochondrial
DNA (mtDNA), synthesis of mitochondrial proteins, trans-
port and incorporation of nuclear-encoded proteins into
the corresponding substructures and supercomplexes of
mitochondria, and dynamic fusion and ssion events
(399). Mitochondrial dynamics are essential to maintain
quality and are sensitive to physiological and pathological
stimuli (400). Whereas fusion leads to elongated mito-
chondrial tubular networks (401), ssion induces fragmen-
tation of the mitochondria allowing the sequestration
of damaged or dysfunctional organelles by mitophagy
(discussed in sect. 4.6.2) but also mitochondrial biogene-
sis (402). More specically, ssion at the periphery of the
mitochondria enables the degradation of damaged com-
ponents of the network, whereas ssion at the midzone
of the mitochondria seems to be instrumental for the pro-
liferation of mitochondria, which cannot be generated de
novo (403). The fusion of the outer and inner membrane
of the mitochondria is regulated by the transmembrane
proteins mitofusion 1 (MFN1) and MFN2 and optic atrophy 1
(OPA1), respectively (401), and important proteins involved
in the ssion process include dynamin-related protein 1
(DRP1), mitochondrial ssion 1 (FIS1), and mitochondrial s-
sion factor 1 (MFF1) (399,402). How mitochondrial dynamics
are altered in muscles of highly trained athletes is not com-
pletely understood (404),andmostdataarebasedon
changes in the abundance of proteins regulating ssion
and fusion dynamics (289,405). It has been suggested
that ssion is elevated after acute exercise to promote
the removal of damaged organelles, whereas fusion
events are increased during the recovery phase (399,
400,406), leading to the elongation of mitochondria
(407). A functional mitochondrial reticulum is important for
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1719
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
ATP generation as well as energy distribution within mus-
cle cells (408). In line with these observations, mitochon-
dria of oxidative muscle bers are highly interconnected,
forming substantially larger mitochondrial networks,
whereas glycolytic bers have more fragmented mito-
chondria (407,409). Recently, it has been suggested that
the two distinct subpopulations of mitochondria (subsar-
colemmal and intermyobrillar) are physically connected,
building a large mitochondrial network (410,411). These
interconnected mitochondrial networks enable a rapid
exchange of various factors between mitochondria to ulti-
mately improve ATP production in a spatially coordinated
manner (402). As demonstrated in murine muscles, the
enrichment of complex IV in subsarcolemmal regions
near capillaries facilitates the generation of the proton
motive force in oxygen-rich areas whereas the higher
abundance of complex V within the myober helps to
generate ATP near the site of high energetic demand
(412). Elongated mitochondrial networks are related to
enhanced oxidative metabolism (400), whereas frag-
mented mitochondria are less efcient in generating ATP
(402). This dynamic remodeling of mitochondria could
make a substantial contribution to improved mitochon-
drial quality and function in trained muscle.
Closely linked to changes in mitochondrial dynamics,
mitochondrial biogenesis plays a major role in the adapt-
ive response to endurance training, with mitochondrial
volume density being strongly correlated to _
VO2max
(254,285,413). The exercise-induced pathways
including Ca
21
signaling and metabolic, oxidative,
and heat stress that are instrumental for mitochon-
drial adaptation are discussed in sects. 4.2.1, 4.4,
4.5.1, 4.5.2, and 4.6. In the vastus lateralis muscles of
untrained individuals, mitochondria comprise 45%
ofthemusclevolume,thehighestbeingintypeIand
the lowest in type IIX bers (255,289,300,302,414).
In endurance-trained athletes, mitochondrial volume
density is 50% higher and citrate synthase activity is
elevated by 74% (254,285). Accordingly, mitochon-
drial respiration is substantially enhanced in elite ath-
letes (415). The exercise training-induced increase in
mitochondrial volume density occurs in all ber
types, including IIX, and ranges between 10% and
60% depending on the training impulse (414,416). In
addition to the greater mitochondrial volume density,
accompanied by an absolute increase in crista sur-
face, crista density is also higher in skeletal muscle
from elite athletes (290), resulting in a further eleva-
tion of muscle respiratory capacity. Thus, besides
improving oxygen delivery to the mitochondria,
muscles can enhance respiratory capacity to meet
the high energy demand during exercise by increas-
ing mitochondrial and crista density as well as opti-
mizing the interconnected mitochondrial networks.
3.4.2. Lipid and glycogen storage in muscle.
Skeletal muscle is a major site for both glucose (in the
form of glycogen) and lipid (intramyocellular triglycer-
ides) storage. Endurance, but not strength/power,
training substantially increases the size of these
depots: lipid content is approximately twofold higher in
the trained musculature of endurance athletes, which, in
part, facilitates the higher rates of fatty acid oxidation (417,
418). Although lipid droplet size is similar in muscle from
trained and sedentary individuals, the total intramyocellu-
lar lipid pool is substantially higher in endurance-trained
muscle.Thisisduetothecombinationofanincreased
number of lipid droplets along with a greater proportion
of type I muscle bers that have a greater capacity for
lipid storage than type II bers (417,419,420). These
droplets can be categorized into subsarcolemmal or inter-
myobrillar lipids, and whereas an elevated fraction of
subsarcolemmal lipid droplets is associated with insulin
resistance (421), endurance-trained individuals predomi-
nantly store lipids in the intermyobrillar fraction, often in
close proximity to mitochondria, which favors high turn-
over kinetics and confers insulin sensitivity (417,422), a
phenomenon termed the athletesparadox(423).
Glycogen, a branched glucose polymer, is found in
three distinct subcellular compartments, with the major-
ity (7585%) located between the myobrils (intermyo-
brillar) near the SR and the mitochondria (424). Between
5% and 15% of the glycogen granules are located
beneath the sarcolemma (subsarcolemmal) and 515%
between the contractile laments within the myobrils
(intramyobrillar) (424). Whereas type I bers have more
subsarcolemmal and intramyobrillar glycogen, type II
bers are enriched in intermyobrillar glycogen (424).
The ber type-specic location of glycogen favors the
functional characteristics of type I and II bers (i.e., fa-
tigue resistance and fast contraction time, respectively)
(351). Intermyobrillar glycogen content is associated
with faster relaxation time and, at least in the type II ber,
is necessary to sustain the high rates of ATP turnover by
the SERCA pumps (351). In contrast, intramyobrillar gly-
cogen is correlated with Ca
21
release from the SR and
fatigue resistance (256,425). Accordingly, in type I
bers of endurance-trained athletes, intramyobrillar gly-
cogen content is 6065% greater, and in both ber
types subsarcolemmal and intermyobrillar glycogen is
increased by 6065% and 2025%, respectively, com-
pared with non-athletes (351). During prolonged exercise
to fatigue, muscle glycogen concentration decreases in
all subcellular compartments, with a preferential deple-
tion of intramyobrillar glycogen (underlying molecular
signals are described in sects. 4.1 and 4.5.1). As pre-exer-
cise muscle (and liver) glycogen content is strongly corre-
lated with prolonged (>90 min) submaximal exercise
FURRER ET AL.
1720 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
capacity, higher muscle glycogen in trained muscle plays
an important role for maximal performance (386,424,
426). In addition, glycogen stores are not only replen-
ished more rapidly in trained muscle but also to a
greater extent (supercompensation) compared with
untrained muscles (described in sect. 4.8.3) (427,
428). The restoration of muscle glycogen can be
accelerated by high carbohydrate availability in the
rst 3 h after exercise and can reach rates of synthesis
of >10 mmol/kg wet muscle weight/h (428). Despite
the capacity for high rates of fatty acid oxidation dur-
ing submaximal exercise, when highly trained athletes
compete in endurance events lasting up to 3 h carbo-
hydrate-based, not fat-based, fuels are the predomi-
nant energy substrate for the working muscles.
Accordingly, carbohydrate and not lipid availability
becomes rate limiting for performance in this context
(194).
3.5. Muscle Memory
Skeletal muscle mass, _
VO2max, and other endurance
training-induced adaptations rapidly decline upon ces-
sation of a training stimulus (i.e., detraining, decondition-
ing), with many structural, metabolic, and performance-
related parameters returning to pre-training values
within weeks to months, even in athletes with a lifelong
history of training. However, prior strength training
seems to facilitate the regain of muscle mass, surpass-
ing the gains that were achieved when training was
commenced from the naive state (429). This phenom-
enon is referred to as muscle memory(430,431)and,
in part, is based on motor learning, intra- and intermus-
cular coordination, prior experience of body perception,
resilience to give into pain and fatigue, and anticipation
of exertion. In addition to these central mechanisms,
there is evidence for a cellular memory in muscle bers
(432). According to the myonuclear domain theory, myo-
nuclei should be lost during detraining-induced muscle
atrophy. However, at least in animal models, the number
of myonuclei remains elevated after a short period of
disuse despite a loss in muscle mass (433). Thus, the
greater myonuclear density and high transcriptional
potential could facilitate muscle growth during retrain-
ing, contributing to muscle memory. In humans, this phe-
nomenon has received little scientic enquiry, and it is
not known whether muscle memory exists, or for how
long accrued myonuclei might be preserved (361). A
large interindividual heterogeneity of the number of
myonuclei elevated after detraining has been reported,
and in most cases the pre-training number of myonuclei
is similar to that after detraining (434). Of note, athletes
with a history of abuse of testosterone or other anabolic
steroids, which boost myonuclear accretion, could still
benet from the elevated number of myonuclei even af-
ter doping cessation, leading to a potential unfair
advantage in competitions long after bans have been
served (435).
Although the enhanced growth rates of muscle mass
during retraining suggest the presence of some kind of
muscle memory (430), many of the training-induced
adaptations in muscle are the result of the complex tran-
scriptional response to repeated bouts of exercise,
which necessitates accessible genomic regions for the
transcriptional machinery. An open chromatin state is
generally indicative of enhanced transcriptional activity
(360), with the accessibility of chromatin associated with
modications of nucleotides in the DNA and posttransla-
tional modication of histone proteins (histone code)
(436,437). Transcriptionally silent genes exhibit closed,
condensed chromatin (heterochromatin), an enrichment
of hypermethylated DNA (5-methylcytosine instead of
cytosine), deacetylated histones, and methylation of dis-
tinct histone residues [e.g., histone 3 lysine residue 9
(H3K9) or H3K27]. Gene transcription requires a state of
open chromatin (euchromatin), linked to demethylated
DNA, acetylated histones, and the methylation of
other histone residues (e.g., H3K4 or H3K26). Besides
stable epigenetic markers that can be passed on to the
next generation, DNA methylation also occurs as a
dynamic process and is inuenced by numerous stimuli
including habitual level of physical activity, nutrient avail-
ability, and (psychological) stress (360,435,438).
Accordingly, promoter regions of genes involved in meta-
bolic pathways, myogenic processes, or oxidative stress
responses are hypomethylated and more accessible in
lifelong physically active compared with inactive men
(439). Inversely, the promoter region of the peroxisome
proliferator-activated receptor (PPAR) ccoactivator 1a
(PGC-1a) is hypermethylated, and thus less accessible, af-
ter bedrest and associated with reduced mRNA expres-
sion (440). In recent years, a growing body of evidence
suggests the involvement of altered chromatin landscape
in response to exercise training as a possible contributor
to muscle memory (360,430). For example, after endur-
ance or resistance training, widespread changes in DNA
methylation status have been reported in muscle (438).
Thus, exercise-responsive genes such as PGC-1aare
hypomethylated before their induction in response to an
acute bout of high-intensity endurance exercise (441).
Despite these observations, the contribution of DNA meth-
ylation to training adaptation and muscle memory is con-
troversial. In some studies, endurance training promoted
the demethylation of genes involved in angiogenesis or
oxidative metabolism, associated with increased gene
transcription (442,443), whereas in others there was little
effect of HIIT or resistance training on DNA methylation
(444). A recent study reported that despite divergent
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1721
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
contractile stimuli (HIIT, endurance and resistance exer-
cise), changes in DNA methylation and mRNA expression
in skeletal muscle were largely conned to the late (48h)
recovery period and similar between the different exercise
challenges (445). Many of the discrepancies between
investigations can likely be explained by the timing of the
biopsy (446), with time course studies suggesting that
some DNA methylation changes are retained for up to 48
h after the last training bout (442,443), returning to base-
line levels after 72 h (444). It is not known whether DNA
methylation changes are retained after detraining and
thereby contribute to muscle memory. Finally, the differ-
ence of the transcriptional response of an untrained com-
pared with a trained muscle and the involvement of
epigenetic changes therein are currently unexplored.
In contrast to endurance training, the evidence for
long-term epigenetic remodeling triggered by resistance
training is more robust. In response to a training protocol
of loading, unloading, and reloading, several CpG islands
remained hypomethylated during unloading, likely con-
tributing to the elevated transcriptional response of the
associated genes during reloading (430). These results
suggest that chronic changes in DNA methylation may
contribute to the transcriptional memory (430), although
only a small fraction of the differentially methylated genes
displayed a distinct expression pattern (443,447).
Nevertheless, epigenetic regulation of a few regulatory
genes might be sufcient to induce a faster and/or
greater response to recurring challenges.
Collectively, our understanding of the epigenetic
changes in a trained state and the contribution to mus-
cle memory is rudimentary. Findings pertaining to epige-
netic remodeling in human muscle are heterogeneous
regarding the training protocols employed, the study
design/methodology, and the caliber of subjects under
investigation. Individuals described as low respond-
ersto a training intervention may show attenuated epi-
genetic modication compared with high responders,
although further work is needed to corroborate this hy-
pothesis (448). In elite athletes, data on epigenetic pro-
les are almost exclusively from circulating blood cells
but not skeletal muscle (449,450). However, the pres-
ence of polymorphisms of genes encoding proteins
involved in DNA methylation in elite athletes implies a
possible epigenetic predisposition (451).
3.6. How Are Physiological and Cellular Training
Adaptation Brought About?
Many of the training-induced morphological, biochemical,
physiological, and functional adaptations are the culmina-
tion of long-term (weeks to months) exposure to training
stimuli (FIGURE 6). Many of the transcriptional changes
that underpin these adaptations are the transient effect of
repeated bouts of acute exercise that accumulate over
time and result in new steady-state transcript and protein
levels (87,198), even though disparate outcomes for tran-
script and proteins can occur (452). For example, the tran-
scription of many mitochondrial genes is transiently
induced after a single bout of endurance exercise, leading
to mitochondrial biogenesis, improved mitochondrial func-
tion, and elevated oxidative metabolism when the stimuli
are repeated over time (87,405). This response, how-
ever,isnotuniformacrossalltranscriptsandthepro-
teins they encode. For example, in contrast to the
change in trained muscle, robust transcriptional regu-
lation of myosin heavy chains is not observed after
acute exercise bouts (453), whereas other genes
show an attenuated expression after repeated exer-
cise exposure (438). Few studies have simultaneously
investigated contraction-induced changes in mRNA
levels and subsequent training-induced changes in
protein levels in human skeletal muscles following
chronic interventions, and it is clear that exercise-
induced increases in mRNA levels do not always pre-
cede increases in the proteins they encode (405,452,
454). Clearly, several mechanisms regulate the train-
ing response/adaptation, and in the case of transcrip-
tional networks there may be additive or attenuated
responses over time. In sect. 4 we discuss our current
understanding of the molecular mechanisms that are
involved in the acute response of muscle to endur-
ance or resistance exercise bouts.
4. ACUTE MOLECULAR MECHANISMS
UNDERPINNING ENDURANCE- AND
RESISTANCE-BASED EXERCISE
In this section, the sensors and major signaling path-
ways involved in the response of skeletal muscle to a
single bout of endurance and resistance exercise are
summarized. We review the downstream effects of
these stimuli (e.g., transcriptional regulation and transla-
tional control) that promote muscle adaptations in
response to two distinct training paradigms. In addition
to the well-described pathways such as Ca
21
-depend-
ent pathways, AMP-activated protein kinase (AMPK),
and mammalian target of rapamycin (mTOR) complex 1
(mTORC1) signaling reviewed above, we delineate the
important roles of other transducers such as mechano-
sensing and transduction, for which there is a scarcity of
data on athletic populations. A brief discussion of the
common signaling pathways activated by both endur-
ance and resistance training is provided, followed by a
comprehensive overview of the molecular events that
occur after an acute bout of exercise and underpin the
FURRER ET AL.
1722 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
differential responses to divergent contractile stimuli
described in sects. 2 and 3.
A single bout of exercise sets in motion a complex pro-
gram of interconnected signaling events, along with the
activation of numerous biochemical pathways and tran-
scriptional networks that orchestrate the spatio-temporal
responses to muscle contraction and coordinate a pleio-
tropic response in other tissues and organs to control
energy substrate provisioning, oxygen availability, and
heat dissipation (9). These perturbations are initiated by
numerous inputs, which can occur in parallel, overlap, or
be completely independent. Several critical regulatory
nodesprovide the hub for signal integration and subse-
quent control of transcription and enzymatic activity (9,
55,455).Repeatedboutsofexerciseoverseveralweeks
or months result in a continuous modulation of this
response and, over time, ultimately contribute to provoke
chronic adaptations (456). Although the molecular mecha-
nisms underpinning many of the chronic responses to
exercise training remain undened, numerous insights
regarding acute exercise response have been described
in recent years (9,39,41,55,163,393,455,457). A caveat
is that although the focus of this review pertains to data
obtained from exercise-trained humans, many of our cur-
rent mechanistic insights have originated from rodent and
other in vivo and in vitro model systems. It is important to
make a distinction between voluntary, whole body in vivo
responses to exercise and those elicited by other experi-
mental models. Ex vivo electrical stimulation of an isolated
skeletal muscle, for instance, evokes an action poten-
tial and contractionand triggers intracellular path-
ways with putative roles in training adaptation. However,
whole body, voluntary exercise induces a range of addi-
tional physiological responses that are critical for muscle
performance (and movement). Accordingly, many effects
observed in animals and isolated systems can differ from
those seen in humans in vivo, and care should be taken
when extrapolating responses from one set of conditions
or a given experimental model to another (9). Here, we
describe the signaling pathways and mechanistic events
that are principally involved in the response to an acute
exercise bout and culminate in the subsequent training
adaptation. Mechanisms that are important in muscle at-
rophy and pathological situations are not discussed in
detail, and because of space limitations we cite recent
reviews that serve as starting points for further reading
and collections of primary literature. The following sec-
tions are structured based on the putative engagement
oftherespectivepathwaysinmusclecontraction,from
pre (anticipation)- to peri (start of activity, during the exer-
cise bout)- to post (muscle fatigue and exercise cessation
and nally recovery, repair, and regeneration)-exercise
(FIGURE 9). It should be noted, however, that the exact
temporal sequence of engagement and the interactions
between and integration of these processes are not fully
understood.
4.1. Neuroendocrine Signaling in the Anticipatory
Phase and During Muscle Contraction
Anacuteboutofexerciserepresentsaone-offstressor
to whole body homeostasis, provoking widespread pertur-
bations in numerous cells, tissues, and organs that are
caused by or are a response to the increased metabolic
activity of contracting skeletal muscles (9,55,455).
Induction of the ght or ightresponse, including activa-
tion of the sympathetic nervous system in parallel with the
motor system, responds to feedback from the exercise
pressor reex via group III/IV skeletal muscle afferents
(458,459). This reex encompasses feedback that is
evoked from mechanically (muscle mechanoreex) and
metabolically (muscle metaboreex) sensitive afferents
during contractions, leading to parasympathetic depres-
sion and sympathetic activation (460). Consequently,
blood ow to skeletal muscle is increased as a result of
elevated heart rate, blood pressure, and rate of ventilation.
Theexercisepressorreex can be complemented by
central command,in which stimulation of medullary
and spinal circuitries by higher brain centers likewise
evokes respiratory and cardiovascular modulation (460).
Importantly, the sympathetic nervous system can also be
engaged by anticipation and other emotional factors pre-
ceding motor activation. Exercise leads to a substantial
increase in circulating catecholamines, a response that is
greater in trained compared with untrained individuals
exercising at the same relative intensity (458,459). In part,
centrally controlled modulation of the sympathetic nervous
system is required for a systemic activation of events that
support muscle contraction, such as increased pulmonary
and cardiovascular output or various metabolic pathways
to liberate energy substrates. Importantly, muscle tissue is
also innervated by the sympathetic nervous system
through the activation of b
2
-adrenoreceptors by catechol-
amines, with epinephrine having a higher afnity than nor-
epinephrine for these receptors, and a higher density of
b
2
-adrenoreceptors on type I compared with type II muscle
bers (458,459). Besides affecting the microvasculature,
the adrenergic system exerts other functions in this tissue,
including direct effects on the neuromuscular system
(FIGURE 10)(458,459). Initially, adrenergic action on the
presynaptic side of the NMJ helps synchronize neurotrans-
mitter vesicle fusion and augment acetylcholine release.
Then, b
2
-adrenoreceptor action on the muscle bers acti-
vates the Na
1
-K
1
pump and thereby ber excitability,
potentially attenuating fatigability. Muscle contractility, in
particular twitch force and relaxation rate, is modulated by
the adrenergic effect on Ca
21
release and reuptake via
RYR1 and phospholamban, the latter of which is exclusively
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1723
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
expressed in slow muscle bers. Metabolically, b
2
-adreno-
receptor activation antagonizes insulin to stimulate glyco-
gen breakdown and inhibit glycogen synthesis. Moreover,
b
2
-adrenoreceptor agonism represses proteolytic proc-
esses, leading to a transient anabolic effect on muscle
mass. As members of the G protein-coupled transmem-
brane receptor family, b
2
-adrenoreceptors engage numer-
ous signaling pathways and effectors in muscle cells (461).
The exchange of GDP for GTP at the astimulatory sub-
unit of guanine nucleotide-binding regulatory protein
(Gas)resultsinanactivationofadenylatecyclase,the
cyclic AMP (cAMP) signaling pathway, and ultimately
elevated transcriptional activity of the cAMP response
element-binding protein (CREB), which stimulates the
modulation of additional control genes, including
PGC-1a, or myogenic factors such as myoblast deter-
mination protein 1 (MyoD) (461). The inhibition of the
Forkhead box 3 (FOXO3) by PGC-1a, and hence of pro-
tein degradation and ber atrophy (462), is potentiated
by the effect of the Gbc subunit of the b
2
-adrenoreceptor,
which modulates the activity of phosphoinositide 3-ki-
nase (PI3K) and, via downstream activation of protein
kinase B (PKB/Akt), exerts a negative effect on FOXO3
function as well as a positive modulation of protein syn-
thesis by activation of mTORC1 (461).
4.2. Motor Neuron Activation of Muscle Fiber
Contractions
Muscle ber contraction can be initiated via different
mechanisms: neuronal activity in the motor cortex for
voluntary movement, sensory neuronal input for involun-
tary reex contractions such as elicited in proprioceptive
or vestibular control, or hypothalamic activation of ther-
mogenesis-promoting neurons for shivering (463465).
Regardless of the origin of the signal and upstream cir-
cuitry, neuronal input converges on a-motor neurons in
the ventral horn of the spinal cord (466). These motor
neurons, their descending axons, and the innervated
muscle bers form a motor unit that transforms synaptic
input into muscle contractions (318). Motor units differ in
size, with one motor neuron interfacing with from a
handful to many thousands of individual muscle bers
(318). Collectively, the motor units of one muscle are
Anticipation
Initiation
Activity
Recovery
Exercise
Time
1
1
2
2
3
3
4
4
5
Stress response:
mechanical
redox
heat
proteostatic
energetic
substrate availability
oxygen availability
Motor neuron
signaling
Sympathetic nervous
system activation
Muscle fiber repair and regeneration
Substrate refueling and proteostasis
Functional retrieval
Central and peripheral
processes, e.g., redox signaling
Circadian
rhythms
Fatigue/
Exhaustion
5
FIGURE 9. Temporal engagement of different pathways and processes in skeletal muscle in exercise. Anticipation of activity is linked to elevated
sympathetic nervous system tone. Motor neuron ring triggers muscle contractions at the start of and during an activity. While contracting, muscle will
be affected by different stressors and the related reaction/mitigation, for example, mechanical stress, reactive oxygen and nitrogen species (redox) pro-
duction, heat, altered proteostasis and protein unfolding, metabolic changes, substrate availability, and oxygen provisioning. The exact temporal
sequence of engagement of and interactions between these pathways are unknown. Different mechanisms contribute to muscle fatigue, exhaustion,
and exercise cessation. Subsequently, muscle repair, regeneration, and refueling are important for functional retrieval. Many of these processes are
modulated by circadian input. Images: triathlon anticipation from Wikimeda Commons (CC-BY-SA 3.0, creator: Wiech), triathlon start from Wikimedia
Commons (CC-BY-SA 2.0, creator: IQRemix), triathlon cycling from PxHere.com (CC0 Public Domain), exhaustion from Wikimedia Commons (CC-BY-SA
4.0, creator: Wallco26), massage from Freepick.com (author: javi_indy), waking up from Freepik.com (author: diana.grytsku).
FURRER ET AL.
1724 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
referred to as the motor unit pool (318). At the synapse
between motor neurons and muscle bers, the NMJ, an
axonal action potential, results in the release of the neu-
rotransmitter acetylcholine, which, after traversing the
synaptic cleft, binds to nicotinic acetylcholine receptors
that cluster at the entry points of subneural clefts on the
muscle membrane (FIGURE 11A)(467,468). Activation of
these ligand-gated ion channels results in inux of Na
1
and a process called ECC, leading to depolarization of
the muscle membrane and, on the intracellular side of T
tubules, the release of Ca
21
from the SR, which ultimately
initiates and maintains muscle contractions (469). Thus,
one of the rsteventstooccurinresponsetoasingle
bout of exercise is initiation of Ca
21
-dependent signaling.
4.2.1. Ca
21
signaling.
Besides binding to troponin C and initiating the interac-
tion between actin and myosin, intracellular Ca
21
acti-
vates a multitude of signaling pathways in myobers,
modulating numerous physiological functions (470). For
example, Ca
21
signaling is associated with the control of
glycolysis, mitochondrial function, and the rates of protein
synthesis and degradation (471). The different Ca
21
tran-
sients that are evoked by activation of muscle bers by
slow and fast motor neurons contribute to the specicity
of different muscle ber types (472,473). Mechanistically,
the protein phosphatase calcineurin A (CnA) and Ca
21
/
calmodulin-activated protein kinase II (CaMKII) are Ca
21
-
activated mediators involved in controlling gene tran-
scription linked primarily to a slow ber phenotype (471).
In this context, CnA and CaMK activities converge on the
cAMP-dependent binding protein (CREB) and activating
transcription factor 2 (ATF2, also called CREB2), which
are activated by phosphorylation and dephosphorylation
of respective Ser/Thr sites, subsequently binding to the
promoter of the PGC-1agene PPARGC1A and inducing
transcription of this coregulator protein, among others
(474). Once synthesized, PGC-1acompetes with histone
deacetylase 5 (HDAC5) to coregulate myocyte enhancer
factor 2 (MEF2) family members on its own promoter and,
using this positive autoregulatory loop, ensures robust
transcriptional expression (475,476). Consistent with ele-
vated levels in slow and exercised muscle bers, PGC-1a
mediates a broad remodeling of skeletal muscle, result-
ing in a slow-twitch, oxidative, fatigue-resistant phenotype
(477)thatalsoincludesextramyober adaptations such
as at the microvasculature (478)orthepresynapticside
of the NMJ (315). Mice with a skeletal muscle-specic
ablation of PGC-1adisplay abnormal glucose and insulin
homeostasis (479), contraction-induced ber damage,
impaired endurance capacity, and other characteristics
indicative of pathological inactivity (480).
The mechanisms that underlie the broad integration
of a vast number of signaling pathway that are engaged
in contracting muscle, and which mediate a tightly chor-
eographed modulation of broad transcriptional pro-
grams by PGC-1a, are unclear. Gene expression from
different promoters and transcriptional start sites (474,
481,482), various isoforms (481,482), context-specic
posttranslational modications (474,482), and the RNA
binding-dependent assembly of specicmultiprotein-
containing transcriptional complexes and DNA regula-
tory elements in sequestered nuclear condensates (483)
could all contribute to a coordinated spatio-temporal
control of PGC-1a-mediated network control (482,484).
Indeed, PGC-1afunctionally and/or physically interacts
with multiple transcription factors and coregulators that
affect the exercise phenotype of skeletal muscle in
a dynamic manner (482,485,486). Thereby, spatial
TF
β2 AR
symp. NS
Afferents
RYR1
SERCA PLN
Proteolysis
cAMP
PGC-1α
PI3K-Akt-mTOR
FOXO3
Synchronization of vesicle fusion
Augmented acetylcholine release
Action
potential
Na+/K+ pump
Fiber
excitability
Ca2+
Fiber
contractility
Ca2+
Glycogen
breakdown
FIGURE 10. Neuroendocrine signaling by
the sympathetic nervous system in exer-
cise anticipation and muscle contraction.
Sympathetic activation of the motor neuron
and skeletal muscle cells results in modula-
tion of ber excitability and contractility,
metabolic and proteostatic remodeling,
and the activation of a transcriptional pro-
gram. These events prepare muscle cells for
upcoming contractions and help to maintain
contractile activity upon engagement. b2AR,
b
2
-adrenoreceptors; Akt, protein kinase B;
cAMP, cyclic AMP; FOXO3, forkhead tran-
scription factor O3; mTOR, mammalian target
of rapamycin; PGC-1a, peroxisome prolifera-
tor-activated receptor ccoactivator 1a;PI3K,
phophoinositide 3-kinase; PLN, phospholam-
ban; RYR1, ryanodine receptor 1; SERCA, sar-
coplasmic/endoplasmic reticulum Ca
21
-
ATPase;symp.NS,sympatheticnervoussys-
tem; TF, transcription factor. Image created
with BioRender.com, with permission.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1725
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
specication, and the expression of synaptic genes in
subsynaptic but not extrasynaptic myonuclei (487), or a
temporal specication, such as the activation of cata-
bolic and anabolic pathways like fatty acid b-oxidation
and de novo lipogenesis to increase intramyocellular lip-
ids (488), might be achieved (474,489). Of note, Ca
21
signaling also activates a number of additional transcrip-
tion factors with unclear epistatic relationship to PGC-1a,
including family members of the nuclear receptor 4A
(NR4A) family, of which NR4A2/NURR-1 and NR4A3/
NOR-1 in turn regulate target genes involved in oxidative
metabolism and improved muscle endurance (490,491).
As Ca
21
signaling is activated in response to both en-
durance- and resistance-based exercise, it is still unclear
whether and how this pathway is affected in a training
mode-specic manner. For example, training intensity
and duration may be one of the factors that inuence
Ca
21
uctuations in the muscle and thereby inuence
Ca
21
-mediated signaling.
4.2.2. (Neuro)endocrine factors, exerkines, and
myokines.
Besides neurotransmitter-mediated activation, motor neu-
rons and skeletal muscle bers engage in a bidirectional
cross talk through several secreted factors (FIGURE 11B).
Some of these, such as motor neuron-derived agrin, are
important for the development and stabilization of the
NMJ (468). This synapse, however, also exhibits remark-
able plasticity in the mature state, with exercise providing
TF
β2 AR
symp. NS
Afferents
ECC
NTs
Fiber contraction
CaMK, CnA
CREB
ATF2
PGC-1α
A
B
TF
Lactate IL-13
GDF3
Paracrine
Endocrine
Paracrine
Autocrine
Action
potential Retrograde
signaling
Ca2+
Retrograde
signaling
Endocrine:
Adipo-, Hepato-, Osteokines
Corticosteroids
Testosterone
Growth hormone => IGF-1
...
Myokines
Myobolites
FIGURE 11. Motor neuron signaling and (neuro)endocrine
effectors of contraction. A: motor neuronal signaling triggers
excitation-contraction coupling and thereby evokes a rise in
intramyocellular calcium (Ca
21
), which enables ber contrac-
tions, activates various signaling pathways, and modulates a
transcriptional response, including retrograde feedback to the
motor neuron. Motor neuron activity is modulated by sympa-
thetic tone and includes various neurotrophic factors besides
the neurotransmitter acetylcholine. B: exerkines, originating
from tissues including muscle (myokines), liver (hepatokines),
adipose tissue (adipokines), and bone (osteokines) as well as
other hormones coordinate a systemic response to contractile
activity. Many of these factors exert auto-, para-, and endo-
crine effects. In addition, signals can be propagated by exer-
cise-linked changes in different metabolites (myobolites or
myometabokines). b2AR,b
2
-adrenoreceptors; ATF2, activat-
ing factor 2; CaMK, calcium/calmodulin-dependent protein ki-
nase; CnA, calcineurin A; CREB, cAMP-responsive element
binding protein; ECC, excitation-contraction coupling; GDF3,
growth differentiation factor 3; IGF-1, insulin-like growth factor
1; IL-13, interleukin 13; NTs, neurotrophic factors; PGC-1a,per-
oxisome proliferator-activated receptor ccoactivator 1a;
symp. NS, sympathetic nervous system; TF, transcription fac-
tor. Image created with BioRender.com, with permission.
FURRER ET AL.
1726 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
a strong stimulus to alter morphology and function (312).
Neuregulin-1 is secreted by the motor neuron and acti-
vates signaling pathways in the muscle ber that lead to
the phosphorylation of PGC-1aand the GA-binding pro-
tein B [GABPB, also called nuclear respiratory factor 2b
(NRF-2b)] subunit of the GABP transcription factor com-
plex (487). These local events, which primarily affect sub-
synaptic myonuclei close to the NMJ, could provide the
mechanistic basis of the spatial specication of PGC-1ato
exclusively regulate the transcription of postsynaptic NMJ
genes in these, and not extrasynaptic, nuclei (487).
Intriguingly, active muscle PGC-1atriggers a remodeling
process of the NMJ that extends to the presynaptic side
with altered mitochondrial and synaptic vesicle numbers
in the active zone and a quantal content reecting slow
NMJ function (315). These observations go beyond the
current view of a unidirectional control of muscle ber
type by the respective activity patterns of motor neurons,
in that elevation of muscle PGC-1a,observedinslow-type
or exercised bers, affects motor neuron function, at least
at the NMJ. Neurturin could be a mediator of this PGC-1a-
dependent retrograde signaling, acting on the NMJ (492)
and even on motor neurons in the spinal cord (493).
Another example of such a factor, brain-derived neurotro-
phic factor (BDNF), a signaling factor secreted from neu-
rons but also cells and tissues such as skeletal muscle
bers (494), functions as a myokine, a hormonal entity
produced and secreted by muscle cells. BDNF expres-
sion is elevated upon contractile activity, the protein
secreted, and besides a potential effect on the motor
neuron affects NMJ morphology and function postsynap-
tically in an autocrine manner (495).
Several other (neuro)endocrine factors have also been
identied in the context of exercise. For many of these,
particularly testosterone and other classical steroid and
non-steroid hormones, regulation during an acute exer-
cise bout is probably of little signicance compared with
the chronic effects such as restoration of substrate stores,
muscle repair and regeneration, and in extreme situations
overtraining/overreaching. For example, corticosteroids,
glucagon, and leptin are elevated when blood glucose
and/or muscle and liver glycogen concentrations are low
and stimulate fatty acid oxidation in muscle by activation
of AMPK (496). Testosterone, growth hormone, and the
downstream target insulin-like growth factor 1 (IGF-1) are
induced after different types of resistance training and
exert anabolic effects by stimulating MPS and ber repair
(497). The regulation of these hormones in humans is
variable, and little is known about the exact molecular
mechanisms that mediate these adaptations. Indeed,
gain- and loss-of-function models of the receptors of
some of these hormones fail to reveal a clear picture
regarding their function in exercise-induced muscle plas-
ticity (486). Besides these classic hormones, a novel class
of so-called myokineshave been discovered and stud-
ied in recent years (498). Myokines, some of which are
only produced in contracting muscle bers, sometimes
referred to as exerkines,elicit auto-, para-, and endo-
crine effects to coordinate local and systemic processes
(499,500).Forexample,whensecretedasamyokine,
interleukin (IL)-6 acts as a metabolic coordinator by pro-
moting lipolysis in adipose tissue, gluconeogenesis in the
liver, and, via activation of AMPK, glucose uptake and fatty
acid oxidation in muscle (501). Paracrine effects of myokines
are also instrumental for an adequate response of muscle
tissue to exercise, such as the proangiogenic effects of IL-6
or the vascular epithelial growth factor (VEGF) on epithelial
cells (501) or the activation of resident macrophage polariza-
tion by the B-type natriuretic peptide (BNP) and secreted
phosphoprotein 1 (SPP1) (502,503). Many of the endurance
exercise-induced myokines are under the transcriptional
control of PGC-1a(501,504), whereas the regulation of
those modulated by resistance training is less clear (504,
505). Paracrine signaling from different cell types to myob-
ers is also important for a normal exercise response, such
as the exercise-induced secretion of IL-13 by type 2 innate
lymphoid and other immune cells that promotes an oxida-
tive phenotype in muscle (506) or growth differentiation fac-
tor 3 (GDF3) by macrophages (507), as well as endothelial
cell-secreted lactate, both important for muscle regenera-
tion (508). A complex dialogue between muscle and other
cell types, mediated by myokines, hepatokines, adipokines,
and osteokines, ensures proper and coordinated local as
well as systemic adaptations to contractile activity, such as
for the adiponectin-mediated activation of muscle AMPK
(509,510). Collectively, however, our understanding regard-
ing hormones that affect skeletal muscle during contraction
or throughout recovery and regeneration from exercise is
rudimentary.
4.3. Mechanosensing and Mechanostress
Mitigation
4.3.1. Cell membrane mechanosensing.
Mechanical stress is exerted on muscle bers at the initia-
tion of and during exercise by passive stretching, sarco-
meric contraction, and other stimuli (FIGURE 12)(511). The
force-induced stretch and contraction/compression of mus-
cle bers in situ is not restricted to the longitudinal direction
but extends across orthogonal, radial, and tangential axes
(511). The ensuing shear, tension, and compressional stress
and cellular deformation present a high potential for dam-
age to the extracellular matrix (ECM), cell membrane, intra-
cellular scaffolds, and other structures. Therefore, a
complex system of mechanosensing exists to attenuate
damage and initiate adaptive processes that confer protec-
tion against acute and subsequent insults. Broadly,
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1727
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
mechanical stress is detected by different structures at the
cell membrane or by intracellular sensors. Stretch-activated
channels (SACs) belong to the rst category in skeletal
muscle, of which the mechano-gated Ca
21
transient recep-
tor potential channels (TRPCs) are activated by mechanical
stress (512). The expression and function of the mechano-
sensing cation channel Piezo1 in skeletal muscle is less
understood, at least in intact bers (513). TRPC activity
results in Ca
21
inux, and the ensuing elevations in intra-
myocellular Ca
21
either directly affect target structures or,
through the activation of distinct signaling pathways, result
in the activation of calmodulin, CnA, MEF2, and nuclear
factor of activated T cells (NFAT) (512). Dephosphorylation
of NFAT by CnA enables a cytoplasmic-nuclear transloca-
tion and subsequent regulation of gene expression by this
transcription factor, often together with MEF2 (512). NFAT
binds to regulatory genes involved in muscle ber type
remodeling or hypertrophy, including induction of PGC-1a
and, via a positive feedback loop, transcription of TRPCs,
thereby sensitizing muscle bers to future mechanoinsults
(512). Non-genomic effects of intracellular Ca
21
include
direct modulation of structural proteins such as titin or
actin. Deformation of cytoskeletal structures thus medi-
ates mechanosensing by receptors at the cell membrane,
leading to a regulatory loop that reacts to changes in me-
chanical stress by adapting sensing and cell rigidity (512).
Ca
21
, for example, affects the assembly of actin laments,
thereby driving a cytoskeletal rearrangement that affects
cell stiffness (512). Increases in intracellular Ca
21
also
result in the activation of Ras guanine nucleotide-releasing
factors (GRFs), which in turn activate the Ras GTPase (514).
In this manner, the mitogen-activated protein kinase
(MAPK) signaling pathway is engaged, leading to elevated
activity of p38 MAPKs, extracellular signal-regulated ki-
nases (ERKs), c-Jun NH
2
-terminal protein kinases (JNKs),
and stress-activated protein kinases (SAPKs). The MAPK
pathway integrates TRPC Ca
21
-derived mechanosensing
via transmembrane adhesion receptors such as integrins,
integrin-associated proteins, focal adhesion kinase (FAKs),
and other components of the focal adhesion complex
(512,515). The stress-activated kinases then execute the
phosphorylation of enzymatic effectors and transcriptional
regulators including PGC-1a(516), serum-response factor
(SRF), JUN, and FOS, forming the activating protein 1 (AP-1)
complex or early growth response gene 1 (EGR-1), and
thereby inducing the expression of immediate-early and
delayed primary response genes (517). This program
includes several transcription factors needed for second-
ary response gene transcription, some of which are
directly modied by stress kinase phosphorylation (518).
Collectively, these early primary and secondary response
genes promote processes related to cell survival, cytos-
keletal rearrangement, and elevation of small heat shock
proteins (HSPs) and other chaperones, thus mitigating
potential harmful events in the contracting muscle ber on
different levels (511,512,518).
4.3.2. Cytosolic mechanosensing.
Besides these mechanosensing systems originating at
the cell membrane, at least two cytosolic signaling path-
ways are important. First, mTORC1 activity is increased
Shear stress Contraction-Relaxation
TRPC
Cytoskeleton
Titin
Hippo
mTORC1
YAP/TEAD
Integrins
FAC
Change in titin stiffness
Repair of sarcomeres
Ca2+
Stress kinases: p38 MAPK,
ERK, JNK, SAPK
Nuclear
deformation
Immediate-early and delayed
primary gene expression:
SRF, EGR-1, AP-1, PGC-1α,...
TF
FIGURE 12. Mechanosensing and mechanostress mitigation in the contracting muscle ber. Mechanical stress is sensed and translated by structures
at the cell membrane and intramyocellular components in the cytosol, cytoskeleton, sarcomeres, or nucleus. As a consequence, resistance to shear
stress is increased, stiffness and integrity of sarcomeric structures adapted, and a broad program of immediate-early and delayed primary genes initi-
ated. AP-1, activating protein 1; EGR-1, early growth response gene 1; ERK, extracellular signal-regulated kinase; FAC, focal adhesion complex; JNK, c-
Jun NH
2
-terminal protein kinase; mTORC1, mammalian target of rapamycin complex 1; p38 MAPK, p38 mitogen-activated protein kinase; PGC-1a,per-
oxisome proliferator-activated receptor ccoactivator 1a; SAPK, stress-activated protein kinase; SRF, serum response factor; TEAD, TEA domain tran-
scription factor; TF, transcription factor; TRPC, mechano-gated Ca
21
transient receptor potential channels; YAP, yes-associated protein. Image created
with BioRender.com, with permission.
FURRER ET AL.
1728 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
in mechanically stimulated bers (519,520). This protein
kinase is the major regulator of cellular protein synthesis
and one of the key inhibitors of autophagy, thereby con-
trolling cell size (521). In prohypertrophic states, mTORC1
is activated and contributes to the regulation of muscle
mass (522) in a controlled manner. Dysregulation of mus-
cle mTORC1 through ablation of the upstream inhibitor
tuberous sclerosis complex (TSC) results in an atrophic
phenotype (523).Incontrast,inducible,skeletalmuscle-
specic ablation of raptor, an essential component of
mTORC1, has little effect on muscle mass in sedentary
mice (524). These ndings suggest that mTORC1 is an im-
portant, but not the only, regulatory factor governing skel-
etal muscle hypertrophy (525). The prototypical activators
of mTORC1 signaling are insulin and amino acids, in partic-
ular leucine, arginine, and glutamine, collectively repre-
senting an anabolic context with high nutrient availability
(521). However, in muscle contraction and mechanosens-
ing, other stimuli may be more important to activate
mTORC1, although the specic exercise-stimulated mech-
anisms have not been identied (519). Candidate path-
ways include IGF-1 signaling or that of the muscle-specic
IGF-1 splice variant mechano growth factor (MGF), whose
activity is increased by mechanical stimulation and could
engage mTORC1 via PI3K-Akt-dependent pathways.
However, even though IGF-1 exerts clear anabolic effects
in muscle, experimental evidence indicates that acute
mechanosensing and activation of mTORC1 might be in-
dependent of PI3K and Akt (519,520). mTORC1 activation
could also be achieved by cross talk with established
mechanosensing pathways such as intracellular Ca
21
engaging mTORC1 via activation of CaMK kinase a
(CaMKKa) or phosphorylation of TSC and raptor by ERK ki-
nases (519,520). Furthermore, different membrane-bound
phospholipases and diacylglycerol kinase f(DGKf)are
activated by stretch, in part via Ca
21
, and then exert vari-
ous functions related to the mechanoresponse (519).
Phospholipases activate the Raf-ERK as well as PI3K-Akt
pathways, thereby regulating two potential upstream regu-
lators of mTORC1 (519). Moreover, phosphatidic acid pro-
duced by phospholipase D or DGKfdirectly stimulates
mTORC1 activity, most likely by binding to the 12-kDa
FK506-binding protein (FKBP12)-rapamycin binding (FRB)
domain of mTOR (519). Then, a mechanosensitive integ-
rin-FAK-TSC2-Ras homolog enriched in brain protein
(RHEB) axis could also converge on mTORC1 (515).
Finally, several other mechanisms have been pro-
posed to mediate the activation of mTORC1 upon me-
chanical stress and loading, including translocation of
TSC2, the cellular redox state, in particular reactive
nitrogen species (RNS), and amino acid availability.
Regardless of the mode of upstream control, mTORC1
activity triggered by loading-induced mechanical
stress results in an upregulation of protein synthesis
and other hypertrophic programs and, in combination
with PI3K-Akt activity, a reduction in catabolic proc-
esses and apoptosis (519).
In parallel to mTOR signaling, the Hippo pathway also
contributes to mechanostress-induced muscle cell remod-
eling (526). The transcriptional coactivator Yes-associated
protein (YAP) is controlled by the Hippo pathway and is
activated upon deformation or rearrangement of actin
and the subsequent regulation of various actin-associated
proteins (526). After cytoplasmic-nuclear translocation,
YAP coactivates TEA domain (TEAD) transcription factor
family members to control the expression of genes linked
to decrease in apoptosis and increase in muscle hypertro-
phy (526). Even though the prohypertrophic function of
YAP is mTOR independent, the Hippo/YAP pathway may
engage in cross talk with the Akt/mTOR pathway to pro-
mote muscle mass gains and with the transforming
growth factor-b(TGF-b)/small worm phenotype/Mothers
Against Decapentaplegic (SMAD) pathway to coordinate
repair and regeneration, respectively (526).
4.3.3. Structural mechanosensing at the
cytoskeleton, sarcomeres, and cell nucleus.
Intramyocellular structural components also contribute
to mechanosensing, mostly lamentous actin and the
sarcomeric proteins myosin and, in particular, titin, a
giant scaffold protein that is essential for forming the sar-
comeric structure (527529). Because of reversible
extension of specic domains, titin acts as a molecular
spring and determines the passive elastic properties of
sarcomeres and muscle bers and can even contribute
to active force and tension generation during muscle
contractions (530). During stretch, physical deforma-
tion of titin leads to the release of several ankyrin-
repeat proteins including cardiac ankyrin-repeat pro-
tein (CARP), diabetes-related ankyrin-repeat protein
(DARP), and ankyrin-repeat-domain protein-2 (ANKRD2)
that act as transcription factors to initiate sarcomerogene-
sis and ber repair leading to adaptive muscle remodel-
ing and increased expression of structural proteins,
including titin, desmin, and dystrophin (527,528). At that
time, a complex formed by the muscle LIM protein (MLP)
is released from titin and interacts with NFAT signaling to
induce prohypertrophic genes. The titin-cap (T-CAP) acti-
vatestheE3-ubiquitinligasemousedoubleminute2hom-
olog (MDM2), which, together with the titin-associated
Ca
21
-dependent cysteine proteases calpain1 and calpain3
as well as the neighbor-of-BRCA1-gene-1 (NRB-1), promotes
protein degradation and autophagy, hence a robust sys-
tem of protein quality control to repair defective sarcomere
structures (527,528). In the active muscle, in addition to
constituting one of the major signaling hubs for mechano-
sensing, titin represents a tunable spring, acquiring
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1729
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
context-dependent stiffness and thereby conferring
enhanced tension and an increased passive force to mus-
cle cells (527,528). Different posttranslational modica-
tions of the titin protein and direct binding of Ca
21
affect
the stiffness of this protein to improve sarcomere integrity
during exercise and promote the effectiveness of acceler-
ated contraction-relaxation cycles (527,528).
Other intracellular structural components that are
involved in mechanosensing and transduction include the
cell nucleus (531,532), often the most stiff element of the
cell (533). The perinuclear cytoskeleton transmits forces to
the Linker of Nucleoskeleton and Cytoskeleton (LINC)
complex and various proteins of the nuclear envelope
such as lamins and emerin. The LINC complex provides a
mechanical connection between the cyto- and nucleoske-
leton. Remodeling of the nuclear lamina by mechanical
force can directly affect gene expression by changing the
chromatin condensation state, histone modications,
gene repositioning, and the binding of transcription factors
to accessible DNA regions. Inversely, the chromatin state
and histone modications might affect nuclear rigidity.
Thereby, a bidirectional link between nuclear mechano-
sensing and transcriptional regulation is achieved (532,
533). How this is affected by exercise is unclear. Signaling
induced by mechanosensing occurs as soon as the mus-
cle contracts, but it is not known whether resistance- and
endurance-type exercise affect mechanosensing in a simi-
lar fashion and thereby activate the same pathways.
Furthermore, it is possible that specic interventions or
types of contractions (i.e., eccentric vs. concentric contrac-
tions) may enhance stretch or shear stress and thereby
promote mechanosensing-induced signaling pathways.
4.4. Oxidative Stress
Oxidative stress and the associated cellular responses
are closely linked to mechanical stress and mechano-
sensing (FIGURE 13)(534). In contracting skeletal mus-
cle cells, reactive oxygen species (ROS) are produced
by NADPH oxidase enzymes NOX2 and NOX4 and to a
lesser extent by mitochondria. Whereas NOX4 seems to
be a more constitutive enzyme important for basal ROS
production in muscle, NOX2 activity is highly regulated
by specic agonists such as angiotensin II or various
cytokines. At the onset of contraction, phospholipase A
2
is activated by elevated intracellular Ca
21
to produce ar-
achidonic acid from the cleavage of membrane phos-
pholipids. Arachidonic acid in turn increases ROS
production by NOX2 and mitochondria (534). Muscle
stretch, compression, or osmotic shock thereby results
in a rapid burst of intramyocellular ROS, which then
serves as physiological signaling agent (535). In addi-
tion, NOX2 activity might be directly modulated by
mechanosensing of the intracellular microtubule net-
work (536). NOX2-derived ROS in turn affect TRPCs
and, in a positive feedback loop, further sensitize muscle
bers to stretch. In a specic range of concentration, ROS
are important to stimulate maximal force generation,
even though the mechanistic aspects of this function are
unclear. RNS could trigger adaptations similar to ROS in
skeletal muscle. Nitric oxide (NO) is primarily produced by
thetypeIneuronalnitricoxidesynthase(nNOS)enzyme
in the contracting muscle ber (537,538). Subsequently,
NO affects several systems in myobers, either by direct
nitrosylation or by activation of NO-dependent guanylate
cyclases and the ensuing increase in cGMP (536539).
NO has a broad-ranging impact on muscle cell metabo-
lism by increasing glucose uptake or reducing the activ-
ities of creatine kinase and several glycolytic enzymes
(536539). Together with NO produced by epithelial
(eNOS) and inducible (iNOS) NOS, positive effects on va-
sodilation and the activation of satellite cells are
achieved. Finally, both ROS and RNS engage numerous
signaling pathways and transcriptional regulators to pro-
mote a myocellular remodeling (540). MAPK signaling
pathways are redox sensitive, and an activation of JNK,
p38 MAPK, and ERK is observed upon elevated levels of
ROS and RNS (537). In terms of transcriptional regulation,
release of the redox-sensitive Kelch-like ECH-associated
protein 1 (KEAP1) enables a cytoplasmic-nuclear shuttling
of the transcription factor nuclear factor erythroid 2-related
factor 2 (NRF2, also known as nuclear factor erythroid-
derived 2-like 2, NFE2L2), which then binds to antioxidant
response elements (AREs) on regulatory sites of various
genes to promote a robust antioxidant response (541).
NRF2/NFE2L2 also induces the expression of nuclear re-
spiratory factor 1 (NRF1) and PGC-1aand increases mito-
chondrial biogenesis and substrate oxidation (541). NRF2/
NFE2L2 gene expression is controlled by an AMPK-PGC-
1aaxis, implying a robust mutual regulation of NRF2/
NFE2L2 and PGC-1a,withbothPGC-1aand NRF2/NFE2L2
being activated by MAPK pathway effectors (541).
Furthermore, PGC-1aregulates the transcription of several
genes encoding antioxidant enzymes, at least in part, in
an NRF2/NFE2L2-dependent manner (541). PGC-1agene
transcription is also controlled by NO via the activation of
cGMP-associated signaling, potentially in a bimodal fash-
ion (541). Upon redox-dependent oxidation of cysteine res-
idues (540) and a resulting stabilization of the tumor
suppressor gene p53, this transcription factor binds to and
activates the promoter of PGC-1atomediatesomeofits
protective effects in muscle during exercise (542). Nuclear
factor κB(NF-κB) and AP-1 are two additional transcription
factors that can be activated in a redox-specic manner
(537). In this context, NF-κB is a strong regulator of antioxi-
dant genes, whereas AP-1 induces a more general stress
response (537). Of note, both transcription factors interact
FURRER ET AL.
1730 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
with PGC-1a: AP-1, together with MEF2 family factors, con-
trols PGC-1agene expression, whereas PGC-1abinds to
AP-1 to regulate some of its target genes (474,537). In
contrast, a mutual negative interaction between NF-κB
and PGC-1amayhelptoregulatehomeostasis
between metabolism and inammation in muscle
(543). The cellular redox state also affects PI3K-Akt-
mTOR signaling to control rates of protein synthesis
and muscle hypertrophy (534,539). First, ROS-medi-
ated activation of Akt results in an increase in mTORC1
function (534), whereas NO exerts a negative effect on
this pathway, presumably to keep hypertrophy in
check (539). Both redox modalities merge in the pro-
duction of peroxynitrite, formed by the reaction of
superoxide with NO, which leads to an activation of
mTORC1 (534). Thus, the redox balance modulates and
ne-tunes muscle hypertrophy. Collectively, adequate
regulation of ROS and RNS is instrumental to react to
mechanical and other types of stress during exercise
and acutely affects contractility (534). Furthermore,
such a contraction-linked increase in ROS and RNS trig-
gers a hormetic response, in which an upregulation of mi-
tochondrial uncoupling proteins, antioxidant enzymes,
and compounds blunts potential damage by future insults
(534). Accordingly, higher levels of antioxidant enzymes
such as superoxide dismutase (SOD) 1 and 2, catalase
(CAT), and glutathione peroxidase (GPx), as well as non-
enzymatic antioxidants such as reduced glutathione
(GSH) that ameliorate redox homeostasis, are observed in
the muscles of trained individuals (544547). Based on
the role of these processes, inhibition of the production of
ROS and RNS with pharmacological and/or nutritional
antioxidants could be detrimental in promoting an optimal
cellular environment to maximize training adaptation
(548). Even though excessive ROS production has been
linked to tumorigenesis, cardiovascular diseases, hyper-
tension, neurodegenerative disorders, and other chronic
pathologies, exercise training substantially reduces the
risk of these diseases, despite the contraction-induced
acute elevations of ROS and RNS in skeletal muscle (534).
These epidemiological observations highlight the impor-
tance of a well-balanced and coordinated redox produc-
tion and detoxication system in muscle, without any
apparent pathological consequences in muscle tissue or
beyond (549).
4.5. Energy Homeostasis, Substrate and Oxygen
Sensing
4.5.1. Energy homeostasis and energetic stress.
Muscle cell contractions are strongly linked to major
metabolic remodeling (FIGURE 14A). In many situations,
Shear stress Contraction-Relaxation
TRPC
NOX
ROS NOS
RNS Glucose
Vasodilation
mTORC1
NRF2/NFE2L2
cGMP
PGC-1α
Stimulation of maximal
force generation
Enhanced stretch
sensitization
Integrins
FAC
Antioxidant and stress
response genes
Satellite cell
activation
TF
FIGURE 13. Redox stress by reactive oxygen and nitrogen species in muscle contraction. Redox regulation of muscle contractility and stress
response by reactive oxygen (ROS) and nitrogen (RNS) species during muscle contraction. ROS and RNS are primarily produced by enzymes at the
muscle cell membrane. Subsequently, a number of downstream effects are promoted, including an increase in energy substrate and oxygen availabil-
ity, enhancement of force generation, improvement of the resilience against oxidative stress, and modulation of a transcriptional program for muscle
remodeling. cGMP, cyclic guanylate monophosphate; FAC, focal adhesion complex; mTORC1, mammalian target of rapamycin complex 1; NOS, nitric
oxide synthase; NOX, NADPH oxidase; NRF2/NFE2L2, nuclear factor erythroid-derived 2-like 2; PGC-1a, peroxisome proliferator-activated receptor c
coactivator 1a; TF, transcription factor; TRPC, mechano-gated Ca
21
transient receptor potential channels. Image created with BioRender.com, with
permission.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1731
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
the demand for ATP as a rapid energy source is only
partially compensated by creatine phosphate system
transfer and the subsequent increase in oxygen-inde-
pendent anaerobic and aerobic metabolism of glucose,
lipids, lactate, ketone bodies, and other energy sub-
strates. This leads to a shift in the relative concentrations
of ATP, ADP, and AMP (393). During recovery from strenu-
ous exercise, and after optimal refueling, this catabolic
energetic state will pivot into an anabolic state and muscle
and liver glycogen as well as intramyocellular lipid stores
will be replenished. In parallel to the shift in adenosine
species, the relative abundance of nicotinamide adenine
dinucleotide (NAD
1
) and NADH as well as other cofactors
involved in myocellular redox reactions is affected by mi-
tochondrial OXPHOS, lactate dehydrogenase activity, as
wellasotherprocesses(550,551). These fundamental
metabolic changes triggered by the increased energetic
demand of contracting muscle cells result in a complex
ATP ADP / AMP
NADH
FAD
Protein
Degradation Synthesis
AMPK
Autophagy
mTOR
SIRT1
PGC-1α
P
Glucose uptake
Glucose
PGC-1α
A
B
Protein
Degradation Synthesis
AMPK
Autophagy
Amino acids
mTOR
Glucose
Fatty acids
Mondo
Metabolites
Glycogen
C
Acute, severe hypoxia/reduced physoxia
HIF-1α
HIF-1α
Prolonged hypoxia/reduced physoxia
HIF-2α
HIF-2α
Mitochondrial
biogenesis and
function
Fatty acid
transport and
oxidation
Ketone body
metabolism
Lactate
metabolism
FADH2
NAD+
Metabolic
remodeling:
Transcription of
fatty acid
oxidation genes,
glucose uptake
and metabolism,
mitochondrial
genes,...
Modulation of
proteostasis
PPARα
PPARβ/δ
Amino acids
Insulin
IGF-1
O2
PGC-1α
ERRα
O2
Reduced PGC-1α and oxidative metabolism
Inhibition of mTORC1 and anabolic processes
Increased glycolysis and vascularization
Increased myoglobin
Increased vascularization (HIF-1α independent)
SIRT1 TF
TF
HIF-1α
FIGURE 14. Energy homeostasis, substrate and oxygen
signaling. A: metabolic stress signaling in muscle contrac-
tion to increase energy provisioning. A decrease in energy
substrate availability leads to the activation of energy sen-
sors that reduce anabolic processes consuming ATP and
induce catabolic pathways to produce more ATP. This
broad response comprises direct modulation of protein
and enzymatic activities by posttranslational modication
as well as the control of a broad transcriptional program.
B: energy substrate signaling senses substrate levels and
leads to metabolic partitioning. Thereby, muscle cell me-
tabolism is coordinated with substrate availability and
anabolism balanced with catabolism. C: reduced oxygen
availability in skeletal muscle is sensed by hypoxia-induci-
ble factor (HIF)-1aand HIF-2ain acute and chronic set-
tings, respectively. HIF-1arapidly reduces pathways that
consume O
2
, while promoting anaerobic glycolysis to gen-
erate ATP. HIF-2apromotes muscle oxygen extraction
and provisioning. AMPK, AMP-activated protein kinase;
ERRa, estrogen-related receptor a;FAD:avin adenine di-
nucleotide; IGF-1, insulin-like growth factor 1; mTOR, mam-
malian target of rapamycin; NAD, nicotinamide adenine
dinucleotide; PGC-1a, peroxisome proliferator-activated re-
ceptor ccoactivator 1a; PPARa/-b/d, peroxisome prolifera-
tor-activated receptor a/-b/d;SIRT1,sirtuin1;TF,transcription
factor. Image created with BioRender.com, with permission.
FURRER ET AL.
1732 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
modulation of signaling cascades and biochemical path-
ways that are crucial for training adaptation. For example,
a shift in the ATP-to-AMP ratio toward AMP leads to the
activation of a signal transduction pathway centered on
AMPK, which in turn phosphorylates downstream protein
kinase substrates (552,553). In addition, AMPK activity
is modulated by the cellular environment, in particular
upstream kinases liver kinase B1 (LKB1) in exercise of short
duration (i.e., seconds to minutes) and CaMKKbduring
prolonged exercise bouts, or by signaling induced by vari-
ous myokines such as IL-6, IL-15, BDNF, and leukemia in-
hibitory factor (LIF) (498,501,554). Although the precise
role of contraction-induced AMPK signaling in the regula-
tion of glucose uptake and fatty acid oxidation remains
controversial (555), AMPK activation leads to a general
catabolic response upon cessation of exercise, including
augmented uptake of glucose and fatty acids, reduced
rates of glycogen synthesis, elevated glycolysis, and
improved mitochondrial activity (554). Furthermore, AMPK
is a potent activator of autophagy and promotes protein
degradation via the ATP-dependent ubiquitin-proteasome
system to liberate amino acids for energy production in
muscle and gluconeogenesis in the liver (554). The regu-
lation of these two processes is tightly coordinated, with
inhibition or activation largely determined by the relative
activity of AMPK and mTOR and the balance between the
catabolic and anabolic state of the muscle cell (556). For
example, ketoacids resulting from the metabolism of the
branched-chain amino acids valine, leucine, and isoleu-
cine can be used to generate ATP via the tricarboxylic
acid (TCA) cycle in a catabolic state, whereas branched-
chain amino acids can also activate mTOR in the face of
high energy availability (557). In the former setting, AMPK
inhibits mTOR activity by phosphorylation of the mTORC1
upstream inhibitors TSC1 and 2 as well as the mTORC1
component raptor (521). Autophagy is directly stimulated
by AMPK-mediated phosphorylation of the Unc-51-like ki-
nase 1 (ULK1) (521,554), whereas protein degradation is
boosted by phosphorylation and thereby activation of
FOXO3 (554), as well as transcriptional activation of
FOXO1 and FOXO3 (558). In an anabolic context, insulin
or amino acid signaling as well as lysosomal recruitment
of mTOR strongly reduce autophagy and catabolic path-
ways so that protein synthesis and lipogenesis are
enhanced (521). S6 protein kinase (S6K) and eukaryotic
translation initiation factor 4E (eIF4E) are two important
mTORC1 phosphorylation substrates that, together with
increased transcriptional activity of RNA polymerases I
and III, subsequently coordinate ribosomal biogenesis
and protein synthesis (521,559). Insulin signaling-depend-
ent activation of Akt results in phosphorylation and nu-
clear exclusion of FOXO1 and 3, and thereby inhibition of
protein degradation, as well as activation of glycogen syn-
thase kinase 3b(GSK3b),whichinturnphosphorylates
eIF2B and b-catenin to promote protein synthesis (560,
561). The strong inhibitory effect of mTORC1 on autophagy
is exerted by phosphorylation of ULK1 and transcription
factor EB (TFEB), a key regulator of lysosomal biogenesis
(521,562). TFEB and TFE3 are activated by Ca
21
/calci-
neurin signaling in contracting skeletal muscle, resulting in
nuclear translocation and activation of autophagy, lysoso-
mal, and mitochondrial gene expression, in part by func-
tionally interacting with PGC-1a(562). Thus, overall, AMPK
and mTORC1 exert opposite functions, promoting cata-
bolic and anabolic processes, respectively, with a direct
inhibition of mTORC1 by AMPK; this reciprocal control is
important to regulate specic adaptations to endurance-
and resistance-based training stimuli but might also be rel-
evant in the context of concurrent training and the training
interference effect (163,521,563). The molecular basis of
the interference effect underlying the compromised
strength gains observed when individuals simultaneously
undertake both endurance and strength training pro-
grams (described in sect. 2) was investigated at the mo-
lecular level by Atherton and colleagues (564), who
studied isolated rat muscles that were electrically stimu-
lated with either low frequency to mimic endurance exer-
cise (3 h at 10 Hz) or high frequency to mimic resistance
training (6 10 repetitions of 3-s bursts at 100 Hz). They
observed selective activation and/or downregulation of
AMPK-PGC-1aor Akt-mTORC1 signaling pathways in
response to these divergent loading patterns. Specically,
they reported that electrical stimulation that mimicked ei-
ther endurance- or resistance-based training switched sig-
naling to either an AMPK-PGC-1a-orAkt-mTOR-selective
state. They termed this activity the AMPK-Akt master
switchand hypothesized that such selective activation of
AMPK-PGC-1aor Akt-mTORC1 signaling could explain
specic adaptive responses to endurance or resistance
training (564). However, Coffey et al. (565)demonstrated
that in highly trained athletes prior training history attenu-
ated the exercise-specicsignaling responses involved
in single-mode adaptations to training and that a degree
of response plasticitywas conserved at opposite ends
of the endurance-resistance adaptation continuum.
Given that genotypes were originally selected to sup-
port diverse physical activity patterns obligatory for
human survival and that modern-day success in
many sporting endeavors requires a high endurance
capacity coupled with superior explosive power, the
conservation of multiple signaling networks to meet
divergent physiological demands seems to make
sound evolutionary and biological sense (1).
AMPK and mTOR can also synergize and converge in
certain cellular contexts (563), potentially representing
the temporal specication of these two proteins in the
response to contractile activity of skeletal muscle, or ad-
aptation to mixed or concurrent forms of training. For
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1733
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
example, after training, AMPK improves insulin sensitivity
of myobers, thereby potentiating the activation of mTOR
(554). Moreover, both AMPK as well as mTOR converge
on the activation of PGC-1ato promote mitochondrial bio-
genesis and oxidative metabolism of energy substrates
(474,521,554). The interaction between these two key
metabolic regulators, and the implications for muscle cell
plasticity, are therefore multifaceted and remain poorly
understood. In addition to mTOR signaling, AMPK also
engages redox-sensitive pathways including the NAD
1
-
dependent protein deacetylase sirtuin 1 (SIRT1), with
AMPK-mediated phosphorylation of PGC-1apreceding
SIRT1-dependent deacetylation and subsequent activa-
tion in skeletal muscle (566,567). Lysine deacetylation by
SIRT1, SIRT3, and other NAD
1
sensorsaswellasthe
counterregulatory effect by acetyltransferases such as
general control non-repressed 5 (GCN5) are not re-
stricted to enhancing and decreasing the activity of PGC-
1a(568) but also other targets, including histones.
Thereby, a tight coupling between the myocellular redox
state and transcriptional regulation in exercise is facili-
tated (569,570). Although the functional interaction
between PGC-1aacetylation and AMPK-dependent phos-
phorylation has been proposed, the cross talk between
phosphorylation events by other kinases, protein methyl-
ation, ubiquitination, sumoylation, or O-linked b-N-acetyl-
glucosamination is less well understood (474,482).
Posttranslational modications of proteins can affect dif-
ferent properties, including stability or turnover, localiza-
tion, interactions with other protein binding partners, DNA
recruitment, enzymatic activity, or structural conformation
(571,572). Indeed, once activated, PGC-1ainteracts with
many different transcription factors, including estrogen-
related receptor a(ERRa), NRF1, NRF2/GABP, MEF2C,
and MEF2D, to coordinate a complex, yet poorly under-
stood transcriptional network encoding the biological
program of endurance exercise adaptation encompass-
ing vascularization, remodeling of the NMJ, or induction
of a slow-type contractile phenotype (474,573,574).
Moreover, PGC-1ainitiates a coordinated transcriptional
network encoding mitochondrial biogenesis and function,
includingTCAandOXPHOS,fattyaciduptake,transport,
and b-oxidation, ketone body and lactate metabolism, as
well as glucose uptake and use in the pentose phosphate
pathway (474,481,575577). Other transcription factors
such as ERRcor PPARb/d, for which the epistatic relation-
ship to PGC-1aand the involvement in the exercise
response are less clear, evoke similar gene programs in
skeletal muscle (485). Finally, a multifaceted interaction
between PGC-1aand other coregulator proteins affects
the activity of PGC-1ain the control of skeletal muscle
plasticity. For example, nuclear corepressor 1 (NCOR1)
competes with PGC-1afor binding to ERRaand PPARb/d,
thereby reducing PGC-1atarget gene expression
mediated by these interactions (486). In addition, trans-
ducers of regulated CREB-binding proteins 1, 2, and 3
[TORC1/2/3, also called cAMP-regulated transcriptional
coactivators (CRTCs)] are coregulators that are strongly
modulated by environmental cues and, in turn, induce
PGC-1agene expression (486). The diversity of interac-
tions and hence the variety in PGC-1a-containing protein
complexes with different coregulators and transcription
factor binding partners likely contribute to the highly orch-
estrated and coordinated transcriptional network control
exerted by this coactivator protein in skeletal muscle in
exercise (474,481486,578).
4.5.2. Signaling mediated by substrates and
metabolites.
Metabolites and energy substrates have additional
effectsonexerciseadaptationinskeletalmuscle,and
even transient and subtle perturbation in cellular homeo-
stasis can trigger broad downstream effects (FIGURE
14B)(84). For example, glycogen binds to a carbohy-
drate-binding module on the AMPKbsubunit to nega-
tively affect AMPK activity (579), providing a molecular
explanation for the enhanced training effect when individ-
uals commence exercise with low muscle glycogen
stores, as in train low(glycogen) protocols (81,579).
Mondo transcription factors are activated by binding of
glucose, leading to a modulation of the expression of
genes involved in glucose homeostasis by MondoA in
muscle, while concomitantly reducing fatty acid oxidation
by PGC-1a, thereby providing an important metabolic
switch between glucose and lipid oxidation in contracting
bers according to the Randle cycle (580). Fatty acids are
ligands for various nuclear receptors, including the
PPARs, and liver X receptors (LXRs), leading to an
increase in rates of lipid oxidation and lipogenesis in skel-
etal muscle, respectively (485). Furthermore, as noted
above, amino acids are potent activators of mTOR and
thus contribute to mTORC1-mediated control of muscle
proteostasis that promotes ber hypertrophy in response
to resistance-based training (522). The requirement for
mTORC1 in inducing exercise hypertrophy seems to
depend on various factors, including temporal aspects or
training stimuli, with considerable mTORC1-independent
contributions (581). Moreover, amino acid supplementa-
tion to activate mTORC1 without concomitant resistance
training is clearly insufcient to induce gains in muscle
mass and strength (582). Succinate, a citrate cell cycle in-
termediate, accumulates in muscle, the interstitial space,
as well as the circulation during exercise. Signaling trig-
gered by succinate binding to succinate receptor 1
(SUNCR1) induces adaptations in the gene expression pro-
grams for axon guidance, neuronal projection, and muscle
regeneration that collectively contribute to endurance
FURRER ET AL.
1734 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
exercise capacity (583). In addition to NAD
1
, other cofac-
tors at the intersection of metabolism and transcription
are likely affected by contractile activity in skeletal muscle,
including avin adenine dinucleotide (FAD) in the TCA
cycle and OXPHOS, a-ketoglutarate in the citrate
cycle, and acyl-CoA and acetyl-CoA in substrate oxi-
dation and the citrate cycle (393). These cofactors are
directly involved in epigenetic gene regulation by
modulating histone and DNA demethylation and his-
tone acetylation (584). Whether such an epigenetic
coupling to metabolism exists in skeletal muscle dur-
ing exercise is unknown, even though epigenetic
mechanisms contribute to exercise-induced muscle
plasticity (578,585). Finally, the extensive metabolic
remodeling in muscle and other tissues during and af-
ter exercise is likely to include many other metabo-
lites with important signaling and regulatory functions
in the training response (586,587), collectively
referred to as myobolites or myometabokines (588,
589). In addition to muscle-intrinsic effects, inter-tis-
sue communication is also mediated by such circulat-
ing metabolites including b-aminoisobutyric acid
(BAIBA)thathelptocoordinatethegeneralsystemic
response (590). In a similar manner, altered muscle
metabolic capacity can affect the circulating levels of
endogenous metabolites and, in the case of aberrant
levels of such metabolites, thereby contribute to a
detoxicationby lowering hyperketonemia (575)or
the conversion of kynurenine into kynurenic acid,
which is unable to cross the blood-brain barrier (591).
One of the rst studies to propose a link between sub-
strate availability and molecular signaling in exercising
human muscle was that of Wojtaszewski and colleagues
(592). They measured muscle signaling responses and
substrate utilization during and after an acute bout of
steady-state cycling in well-trained subjects under condi-
tions in which exercise was commenced with either low or
high muscle glycogen content. After exercise started in a
low-glycogen state AMPKa2andAMPKa1 activity was ele-
vated to a greater extent compared to when the same
exercise bout was commenced with high glycogen con-
tent (592). However, exercise commenced in the lowered
glycogen state was associated with elevated catechol-
amine concentrations compared with the glycogen-loaded
trial, making it difcult to determine whether fuel availabil-
ity and/or humoral factors contributed to the observed
boosted AMPKa2andAMPKa1 activity. Subsequently, the
results of several other investigations demonstrated that,
compared with normal glycogen levels, commencing en-
durance exercise with reduced glycogen availability
increases the phosphorylation of p38 MAPK and transcrip-
tional activation of IL-6, pyruvate dehydrogenase kinase 4
(PDK4), hexokinase, and HSP72 (188). PGC-1amRNA was
also induced to a greater extent (8- vs. 3-fold) after highly
trained cyclists performed a standardized bout of submaxi-
mal endurance exercise with low versus normal glycogen
concentration (191). Even in response to resistance exer-
cise, commencing exercise with low glycogen seems to
promote mitochondrial adaptation, as demonstrated with
increased phosphorylation of p53 and mRNA expression
of PGC-1a(206). Although train low(glycogen) protocols
boost the training response in well-trained athletes, this
training paradigm failed to be superior to conventional pro-
tocols with regard to performance enhancement.
4.5.3. Oxygen sensing.
If oxygen consumption exceeds oxygen availability and
uptake, physiological hypoxia(or reduced physoxia)
occurs, corresponding to a drop of mean oxygen ten-
sion from 30 mmHg to 23 mmHg in contracting skeletal
muscle bers (593595). This drop is already observed
at relatively low exercise intensities, implying that addi-
tional mechanisms might be contributing to oxygen
availability for different structures in the muscle cells,
such as myoglobin function or local intracellular oxygen
levels and gradients (596598). Hypoxia-inducible fac-
tor (HIF)-1aand HIF-2aare the major mediators of
hypoxic stress (FIGURE 14C). Upon reduced oxygen
availability, HIF-1aproteolysis by prolyl-hydroxylases 2
and 3 (PHD2/3) is alleviated and the HIF-1aprotein stabi-
lized (593). HIF-1athen controls gene programs involved
in anaerobic glycolysis to sustain energy production in the
absence of adequate oxygen supply and represses those
programs that are oxygen dependent, such as mitochon-
drial OXPHOS. This is brought about by inhibiting PGC-1a
expression and activity while promoting gene expression
related to vascularization to improve oxygen supply via
the myokine VEGF (593). Furthermore, the hypoxia-re-
sponsive gene DNA damage inducible transcript 4
(DDIT4) encoding the REDD1/RTP801 protein, an activator
of TSC1/2, inhibits mTORC1 and downstream anabolic
pathways, thereby limiting ATP-consuming processes
(599). The paralog HIF-2aplays a permissive role in the
acute hypoxic response compared with HIF-1abut a
greater function in long-term adaptation in which HIF-1a
activity is repressed (593). In this context, PGC-1ainduces
the expression of HIF-2a,ERRa,theAP-1complex,and
VEGF to promote angiogenesis and vascularization in a
HIF-1a-independent manner (474,600). With regard to
training, the temporary hypoxia in muscle can be exacer-
bated by vasoconstriction of capillaries in muscle tissue,
for example by vascular occlusion or peak contraction
(149). Accordingly, the exercise-induced increase in
mRNA expression of all four postulated PGC-1aisoforms
is blunted when endurance exercise is performed with
blood ow restriction (601). However, in the long term,
training under hypoxic conditions may boost the adaptive
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1735
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
response in the muscle. It remains to be determined
whether such training paradigms are more effective in
enhancing the performance of elite athletes than training
under normoxic conditions.
4.6. Thermotolerance, Protein and Organelle
Quality Control
4.6.1. Heat stress.
Contractile activity is linked to the production of heat, with
only a small fraction of chemical energy (25% depending
on the type of muscular activity performed) being con-
verted into external force production (FIGURE 15A)(602).
Elevated muscle temperature has several advantages in
terms of enzyme kinetics and activity and contributes to
vasodilation and increased blood ow, permitting a better
supply of oxygen and energy substrates, as well as more
efcient removal of by-products. Nevertheless, excess
contraction-induced heat production must be dissipated
through vasodilation and sweating. Thermal stress can
result in misfolding of proteins, impairing cellular function.
To mitigate the potentially damaging effects of heat, exer-
cise induces elevated activity and levels of HSPs, most
notably HSP72 of the HSP70 family (603,604). The
upstream mechanisms of HSP activation in exercise are
poorly understood and likely depend on various meta-
bolic, biochemical, and physical factors as well as training
status. Upon heat stress, HSPs are released from binding
to heat responsive factor 1 (HSF1) and act as chaperones
to refold misfolded proteins and control proteasomal deg-
radation and autophagy (603,604). HSF1 in turn promotes
the transcription of PGC-1a, and together these two pro-
teins induce gene expression of HSPs and PGC-1ain an
autoregulatory loop (603,604). This cross talk with
PGC-1amediates heat shock-boosted oxidative me-
tabolism and mitochondrial function (603,604). In
addition to acting as chaperones and inducing an oxi-
dative phenotype, HSPs exert various other effects in
muscle. For example, by inhibiting JNK through direct
interaction and upstream signaling pathways, there is
A
B
Heat
Contractile activity
HSF1
Increased chaperone activity
PGC-1α
HSPs
PGC-1α
Mitochondrial activity
Shear stress
Proteostasis
ATF4, ATF5, CHOP
ATF6α, IRE1α, PERK
PGC-1α
Ca2+ Ca2+ release and reuptake
UPRER
UPRER
UPRmt
Increased mitochondrial
dynamics
Initiation of mito-
and autophagy
Activation of protein
degradation
Inhibition of gene expression
and protein synthesis
Activation of chaperones
for protein refolding
Proteasome
HSF1
TF
FIGURE 15. Heat and proteostatic stress. A:ther-
mosensing and the heat stress response mitigate
misfolding of proteins. Heat is produced by various
processes in the contracting muscle ber, including
shear stress, mitochondrial activity, Ca
21
release
and reuptake, and ATP metabolism in contraction
cycling. Heat is sensed in the cell and a broad tran-
scriptional program engaged to increase mitigating
measures, e.g., chaperones to reduce thermally
induced protein misfolding. B: proteostatic stress
and the ensuing response pathways reduce protein
load, misfolding, and organelle health. Dedicated
pathways in the endoplasmic reticulum and mito-
chondria are engaged by proteostatic dysbalances,
e.g., excessive protein accumulation or misfolding.
At least in part, these two pathways converge to ini-
tiate a transcriptional program aimed at reversing
protein misfolding, alleviating proteostatic stress by
reducing gene expression and protein synthesis
while enhancing protein degradation, and by ensur-
ing organelle functionality. ATF4/5/6a,activating
transcription factor 4/5/6a; CHOP, C/EBP homolo-
gous protein; HSF1, heat shock factor 1; HSP, heat
shock protein; IRE1a, inositol-requiring enzyme 1a;
PERK, RNA-dependent protein kinase-like ER
eukaryotic translation initiation factor 2akinase;
PGC-1a, peroxisome proliferator-activated re-
ceptor ccoactivator 1a; TF, transcription factor;
UPR
ER
, endoplasmatic reticulum unfolded pro-
tein response; UPR
mt
, mitochondrial unfolded pro-
tein response. Image created with BioRender.com,
with permission.
FURRER ET AL.
1736 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
a reduction in inammation and enhanced insulin
sensitivity elicited by HSP72 (603,604). These effects
can be potentiated by heat application to reduce muscle
soreness, improve muscle function during recovery from
damaging exercise, or enhance muscle mass gains
in strength training, potentially triggering a hormetic
response in which the heat shock response is pre-
induced, thereby conferring earlier and/or greater pro-
tection (605). Paradoxically, cold-water immersion
and other modalities to apply cold stressors after
exercise also reduce delayed-onset muscle sore-
ness (DOMS) and inammation after a single applica-
tion (606). However, unlike heat exposure, chronic
cold therapy diminishes strength training effects by
reducing muscle blood ow, attenuating mTORC1
signaling and ribosomal biogenesis, and increasing
FOXO3 activity (606). Repeated cold therapy may
also interfere with the activation of HSPs by exercise
(606). Some current guidelines for the treatment of
soft tissue injuries in sports therefore completely
omit cold application (607).
4.6.2. Endoplasmic reticulum and mitochondrial
unfolded protein response.
HSPs also interact with the regular endoplasmic reticulum
unfolded protein response (UPR
ER
), which is activated by
accumulation of unfolded or misfolded proteins in the ER
(FIGURE 15B)(605). During exercise, interaction between
PGC-1aand cleaved ATF6ainitiates the adaptive UPR
ER
in muscle (608,609). ATF6a, inositol-requiring enzyme 1a
(IRE1a), and the RNA-dependent protein kinase-like ER eu-
karyotic translation initiation factor 2 alpha kinase (PERK)
normally reside in the ER membrane and are engaged
upon ER stress, such as in the context of dysregulated
proteostasis triggered by resistance exercise bouts (608).
Activation drives the release of BiP/glucose-regulating
protein 78 (GRP78), an ER chaperone that then binds to
misfolded ER proteins. IRE1a,PERK,andATF6asubse-
quently initiate signaling cascades aimed at constraining
gene expression and protein synthesis and providing
adequate energy availability to normalize proteostasis
and alleviate ER stress (608,609). Exercise-induced ER
stress is diminished with repeated exercise bouts, indicat-
ing either better control of proteostasis or more efcient
resolution of ER protein misfolding after training (608,
609). Like the ER, upon exercise muscle mitochondria
also initiate an unfolded protein response (UPR
mt
)to
mitigate dysbalanced proteostasis, protein import, and
(re)folding, as well as OXPHOS complex assembly by acti-
vating mitochondrial chaperones and proteases (610). If
overwhelmed, a retrograde signaling is engaged to boost
the activities of the transcription factors ATF4, ATF5, and
C/EBP homologous protein (CHOP), which in turn regulate
the transcription of genes encoding mitochondrial
chaperones and proteases, as well as the integrated
stress response (610). Mitochondrial and cellular pro-
tein quality control are highly coordinated. Initially, cy-
tosolic processes only permit transport-competent
folding congurations of proteins to be imported into
mitochondria, supported by the ubiquitin-proteasome
system that degrades damaged, mislocalized, or
improperly imported proteins at the outer membrane
(611). Then, mitochondrial chaperones and proteases
ensure proper folding and removal of misfolded proteins
within mitochondria (611). The next step in escalation of
damage results in a remodeling of the mitochondrial net-
work by ssion and fusion (402), removal of defective
parts of mitochondrial by mitochondrion-derived vesicles
or piecemeal mitophagy (612,613), ultimately culminating
in wholesome mitophagy (610,614). Mitochondrial dy-
namics, the balance between ssion and fusion, are
instrumental not only for mitochondrial biogenesis but
also for proper morphology of the mitochondrial net-
work (615): fusion leads to larger mitochondria and
increased ATP production, whereas ssion, promoted
in response to severe cellular stress, helps induce the
degradation of dysfunctional mitochondria (402,616).
The cytosolic PTEN-induced kinase 1 (PINK1) is normally
imported into the matrix of healthy mitochondria and
rapidly degraded. Upon deterioration of the membrane
potential or the loss of ATP supply, PINK1 accumulates in
the outer membrane and phosphorylates the E3 ubiqui-
tin ligase Parkin, which in turn ubiquitinates mitochon-
drial proteins to target the organelle for mitophagy. An
acute bout of exercise might lead to elevated organelle
turnover, through both mitophagy as well as general
autophagy, by increasing mitochondrial localization of
Parkin and activation of the AMPK-ULK1 axis, respec-
tively (610,614). However, the regulation and functional
involvement of mito- and autophagy in acute exercise
and chronic training adaptations remain poorly under-
stood (617,618), especially in elite athletes.
4.7. Exhaustion, Fatigue, and Event up to
Cessation of Exercise
The completion of an exercise bout is characterized by
several processes that contribute to exercise cessation.
These can be classied and summarized as exhaustion, a
process dependent on an increasing perception of sub-
jective effort that can, to some extent, be inuenced and
overcome by the individuals willpower and motivation
(619). Physiological pathways resulting in a reduced ability
and ultimately an inability of muscles to contract, inde-
pendent of motivation and willpower, are referred to as
fatigue, in which the continuation of contractile activity
becomes impossible (619). These denitions already hint
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1737
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
B
NOX
ROS
NOS
RNS
RYR1
Reduced fiber
excitability
C
Exhaustion
Fatigue
Decreased motor
neuron firing rate
A
Hyperthermia
Hypoglycemia
Dehydration
Relative O2/CO2 levels
Exerkines/metabolites from
muscle and other tissues
Nociception
Motivation/willpower
Upstream input
Afferent feedback
Performance
ROS/RNS levels
Modulated neurotransmisson
(catecholamine, serotonin,
dopamine, adenosine)
Compromised axonal branch point
propagation (?)
Accumulation of by-products of
contraction and damage
Energy substrate depletion
Oxygen supply
ROS/RNS accumulation
Ca2+ homeostasis dysregulation
Dampened excitation-contraction
coupling
Nociceptor
activation
Decreased
synaptic activity
Na+/K+ pump
Reduced force
production
Desensitization of
proteins to Ca2+
Reduced myosin
ATPase activity
Ca2+
FURRER ET AL.
1738 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
at the multifactorial and multiorgan components that con-
tribute to exercise cessation (FIGURE 16A). Central fac-
tors include catecholamine, serotonin, dopamine, and
adenosine signaling in different regions of the brain
thatevokeastateofmentalfatigue(620). Some of
these alterations in central neurotransmission might
be caused by heat sensing to prevent hyperthermia of
the brain, availability of energy substrates (in particu-
lar hypoglycemia) to prevent catastrophic contractile
failure and inadequate supply for neuronal function,
relative oxygen and CO
2
levels,dehydration,pain/
nociception, or metabolite signaling from contracting
muscle bers such as excess ammonia or increased
ratio of free tryptophan to branched-chain amino acids
in the circulation (619621). Moreover, elevated cyto-
and exerkine proles in exercise could modulate vari-
ous processes, including immune cell activity in the
brain, and thereby also neuroenergetics and neuro-
transmission (622). These, and potentially other cen-
tral mechanisms, most likely help to protect the brain
and other organs from detrimental outcomes trig-
gered by overexertion (623). Central fatigue can be
overcome by direct peripheral stimulation of muscles.
This is not the case for peripheral causes mostly per-
taining to the motor unit, ranging from peripheral nerv-
ous system signaling to the muscle, ECC, energy
supply, and contractility (619). Motor unit fatigue is not
well understood, even though motor neurons are clas-
sied into fast and slow fatigable pools (329,624,
625). Failures in axonal propagation, in particular
across branch points, or in NMJ transmission seem of
little signicanceinexercisesettingsandmightpri-
marily pertain to pathological situations (619). The
observed decrease in motor unit ring patterns and
discharge rates could be mediated by afferent feed-
back and upstream input (619). In some paradigms, fa-
tigue of specic muscles, in particular the respiratory
muscles, will limit performance (626,627). In muscle
cells, different processes have been proposed to
affect contractility and fatiguability, such as contrac-
tion-induced depletion of endogenous fuel stores to-
gether with inadequate provisioning of exogenous
energy substrates and oxygen supply (619,628). The
accumulation of intra- and extracellular by-products
of prolonged contraction or metabolites enriched
because of muscle ber damage could likewise gov-
ern fatigue, such as products of energy metabolism,
oxidative stress, inammation, or respiration (CO
2
)
(619,628). In essence, the mechanistic underpin-
nings of muscle-intrinsic fatigue are not entirely
clear. Interestingly, in contrast to the stimulating
effect of ROS and RNS on muscle contractility at
lower levels, once a critical threshold is reached,
dampening effects of accumulated ROS and RNS on
performance are observed (FIGURE 16B)(629). For
example, ROS induces muscle fatigue, mediated by a
desensitization of receptors and myobrillar proteins
to Ca
21
,andadecreaseinNa
1
-K
1
pump activity
(FIGURE 16C)(534). This concentration-dependent,
biphasic bell-shaped effect on muscle performance
and fatigue might be a protective mechanism against
excessive contractile activity and subsequent tissue
damage, with ROS serving as an internal rheostat for
the strain exerted on bers (534). RNS evoke a simi-
lar response by initiating processes such as muscle
pain and fatigue to mitigate overexertion and over-
load linked to excessive cellular and tissue damage.
For example, NO modulates synaptic transmission
by retrogradely affecting presynaptic structures of
the NMJ, activates nociceptor complexes containing
the NO-sensitive calcitonin gene-related peptide
(CGRP) receptor and thereby causes muscle pain, or
reduces muscle force production, at least in part by
reducing myosin ATPase activity and Ca
21
release
from the SR by affecting the RYR and the SERCA
pumps (536539). Thus, for both ROS and RNS,
modulation of intramyocellular Ca
21
homeostasis is
central for the regulation of fatigue (629). Of note,
the benets of inhibition of the production of ROS
and RNS with pharmacological or nutritional antioxi-
dants to delay fatigue, reduce cellular damage, and
shorten the recovery period are superseded by the
dampening effects on the procontractile and per-
formance aspects of ROS and RNS during exercise,
FIGURE 16. Exhaustion and muscle fatigue. A: central and peripheral contributors to exhaustion (volitional) and fatigue (involuntary). Altered neuro-
transmission in the brain ensures protection of this and other organs from hyperthermia, hypoglycemia, dehydration, shifted O
2
/CO
2
levels, and further
potentially detrimental processes in exercise. To a limited extent, these effects can be overcome by willpower and motivation. Peripheral factors offa-
tigue involve the motor neuron and muscle bers. Although impairments in action potential propagation and neuromuscular junction transmission
seem minor in healthy individuals, input from upstream brain regions and afferent feedback modulate motor neuron ring rate. Muscle-intrinsic contrac-
tility is affected by energy substrate and oxygen availability, accumulation of by-products of contraction and damage, elevation of reactive oxygen
(ROS) and nitrogen (RNS) species, a dampening of excitation-contraction coupling including the Na
1
-K
1
pump, and intramyocellular Ca
21
homeostasis.
B: concentration-dependent, biphasic bell-shaped effect of ROS and RNS on muscle performance. During contractions, ROS and RNS sustain and
enhance contractile activity. However, once levels exceed a poorly dened threshold, a different set of processes is engaged that limits performance.
C: exceeding a certain concentration, ROS and RNS contribute to muscle fatigue by affecting ber excitability and contractility, synaptic activation, and
nociception. A modulation of intramyocellular Ca
21
homeostasis thereby plays a central role. Putatively, this rheostathelps to avoid overexertion and
to minimize muscle tissue damage. NOS, nitric oxide synthase; NOX, NADPH oxidase; RYR1, ryanodine receptor 1. Image created with BioRender.com,
with permission.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1739
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
resulting in an overall reduction in training response/
adaptation (534). Chronic elevation of ROS in inac-
tive bers such as in bedrest leads to pathological
outcomes of constitutively stimulated redox signal-
ing in muscle (630). In this context, restoration of
ROS levels and redox signaling pathway activity by
exercise, pharmacological, or nutritional interventions
might confer clinical benets (630,631). However, antiox-
idant treatment with supplements or nutritional compo-
nents must be carefully tailored to the speciccontext:
oxidative distress and potential antioxidant deciencies
must be conrmed, an evidence-based, personalized
treatment strategy designed, and the outcome of the
treatment monitored (631). Otherwise, ineffective treat-
mentsoradverseeffectsmightresult,forexamplein
blunting oxidative eustress in exercising individuals (631).
4.8. Repair and Regeneration, Multicellular Cross
Talk, and Refueling
4.8.1. Multicellular interactions in muscle repair.
Upon cessation of exercise, a pleiotropic, highly coordi-
nated program of repair, regeneration, and refueling is
initiated. Skeletal muscle tissue comprises a complex
assortment of different cell types, many of which are still
being identied with novel technical approaches such
as single-cell RNA sequencing (scRNA-seq) (632634).
As noted above, it is well accepted that the cross talk
between different cells is involved in exercise adapta-
tion, such as the release of IL-13 from type 2 innate
lymphoid cells (506). To date, however, data describing
the processes involved in repair and regeneration of
muscle tissue have, to a large extent, come from
genetic, pharmacological, and other models of severe
physical damage (635637), with little information
derived from physiological in vivo exercise conditions
(638). Upon damage, a cascade of events is com-
menced in which muscle ber-derived and other signals
activate resident immune cells, promote the inltration
of additional immune cell populations, orchestrate the
activation, polarization, and termination of immune cell
phenotypes, and engage satellite and other myogenic
cells to mediate muscle repair and regeneration
(FIGURE 17A)(635637). In exercised human skeletal
muscle, resident and inltrating neutrophils, leukocytes,
monocytes, and macrophages rst promote an inam-
matory environment, phagocytose damaged tissue, and
clean up cellular debris within hours after an exercise
bout (635637). The inltration and activation of these
immune cells are at least in part orchestrated by a cock-
tail of cyto- and chemokines that is released from myo-
bers and other cells including IL-6, C-X-C motif ligand 8
(CXCL8/IL-8), C-C motif chemokine ligand 2/monocyte
chemotactic protein-1 (CCL2/MCP1), tumor necrosis fac-
tor a(TNF-a), IL-1b, and interferon c(IFN-c)(635637).
However, the composition of this cocktail depends on
the preceding exercise modality, load, and intensity. The
release of a disintegrin, metalloproteinase 8 (ADAM8)
and other proteases and the subsequent remodeling of
the ECM facilitate immune cell inltration into muscle tis-
sue (635637). A shift of macrophage polarization from
the pro-inammatory M1 to the anti-inammatory M2
type is a hallmark for the second phase, in which tissue
is regenerated within hours to days (635637). The
release of IL-10, platelet-derived growth factor (PDGF),
transforming growth factor b(TGF-b), IGF-1, angiopoietin,
VEGF, follistatin, NO, hepatocyte growth factor (HGF),
and other signaling molecules from M2 macrophages, T
cells, mast cells, bro-adipogenic progenitors, and type
2 pericytes boosts brotic activity and other pathways to
remodel the ECM to accommodate repaired and novel
myobers and stimulate the activation of satellite cells
and the formation of capillaries and related proc-
esses that are required to reconstitute muscle tissue
(635637,639,640). Functional retrieval is com-
pleted by the expression of mature myosin heavy
chains and subsynaptic NMJ genes to ensure proper
contraction and innervation (635637). This tightly
coordinated process is critical for proper regenera-
tion and tissue remodeling. The absence or elonga-
tion of the pro-inammatory response or a reduced
anti-inammatory response impairs muscle regenera-
tion and could result in detrimental events such as -
brosis. Regeneration is typically complete within 47
days for most exercise challenges (635637). In
response to muscle-damaging exercise, the inltra-
tion of neutrophils occurs within 24 h, followed by an
increase in macrophages after 27days(637,641).
However, the exact time course of muscle regenera-
tion in response to different exercise paradigms is
poorly understood, and it is unclear how these proc-
esses are engaged in athletes. It remains to be deter-
mined whether trained muscles differ from untrained
muscles in terms of absolute number of specic cells
such as tissue-resident macrophages, lymphocytes,
or bro-adipogenic progenitors or the relative pro-
portion of a certain cell type (i.e., M2 macrophages).
The relative amount of myonuclei compared to total
nuclearnumberislowerinsoleuscomparedwiththe
extensor digitorum longus (EDL) (40% and 60%,
respectively), indicating a higher abundance of non-
muscle mononucleated cells in oxidative muscles that
could potentially contribute to better regeneration
(642). In line with this observation, a trained muscle
might harbor a distinct cellular content and improved
regenerative capacity: for example, mice overex-
pressing muscle PGC-1aexhibit a higher proportion of
FURRER ET AL.
1740 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
A
Auto- and mitophagy
Mitochondrial dynamics
B
C
ATP
ADP
Protein
Degradation Synthesis
mTOR
Amino acids
Insulin
IGF-1
Glycogen
PGC-1α
IMCL
Glucose
Fatty acids
Contractile activity
PGC-1α
Recovery/regeneration
PGC-1α
Glucose-6-phosphateRibose-5-phosphate
Nociceptor
activation
Activation
Polarization Macrophages, neutrophils,
monocytes, FAPs, pericytes, ...
Satellite cells
Stem cell niche
Ca2+
Dysferlin
MG53
Nuclear
migration
Sarcomere repair and
sarcomerogenesis
Altered ratio of
ATP/AMP and
NADH/NAD+
Glucose uptake
Fatty acid oxidation
Mitochondrial activity
De novo lipogenesis
Refueling of
intramyocellular lipid
and glycogen stores
Adequate
energy
availability
TF
TF FIGURE 17. Muscle repair, regeneration, and refu-
eling. A: multicellular interactions and intramyocellu-
lar processes that control muscle repair. Resident
and inltrating cells of different types are engaged
by various signals in a temporally highly orchestrated
manner. Thereby, damaged material is removed,
muscle bers repaired or de novo formed, and func-
tional retrieval achieved. Activation of nociceptor sig-
naling should prevent further exertion and damage
during repair and regeneration. Moreover, the multi-
cellular processes are complemented by intramyocel-
lular pathways to restore membrane and sarcomere
integrity as well as organelle function. B:postexercise
refueling of glycogen, intramyocellular lipids (IMCL),
and protein structures. Depleted intramyocellular
energy substrate stores are replenished after exercise
depending on a systemic, anabolic context, i.e., the
availability of the corresponding substrates and
signaling of anabolic hormones. Because of the ener-
getic demand for protein, glycogen, and IMCL
synthesis, these processes are coordinated with mi-
tochondrial activity and ATP production. C: temporal
specication of catabolic and anabolic processes by
peroxisome proliferator-activated receptor ccoacti-
vator 1a(PGC-1a). To avoid a futile cycle, the cata-
bolic activity of PGC-1aduring contractions must be
separated from the anabolic function in regenera-
tion. The mechanistic underpinnings of the tran-
scriptional specication of this coactivator (and
other regulators with multipurpose roles) remain
unknown. FAP, bro-adipogenic precursor; MG53,
mitsugumin 53; mTOR, mammalian target of rapa-
mycin; TF, transcription factor. Image created with
BioRender.com, with permission.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1741
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
anti-inammatory M2 macrophages in muscle tissue
and show improved regeneration upon muscle injury
(643,644). Whether the time course of muscle regen-
eration is accelerated in athletes with a prolonged
history of training, thereby hastening recovery, is cur-
rently unknown.
4.8.2. Fiber repair and regeneration.
Muscle damage is greater after eccentric/plyometric/
lengthening or isometric exercise undertaken with elon-
gated muscle length compared with concentric/miometric/
shortening contractions (637). Exercise-induced muscle
damage is characterized by a force reduction, systemic
increase in myocellular enzymes and proteins such as cre-
atine kinase (CK) and myoglobin, muscle soreness upon
palpation, but also swelling and a decrease in range of
motion (637). Although some of these symptoms appear
immediately after exercise (e.g., loss of muscle force),
others such as muscle soreness present 2448 h later, of-
ten being disassociated from CK levels, which peak after
34days(637). Different mechanisms exist to deal with
muscle cell damage in an escalating manner (645). For
example, membrane lesions are rapidly sensed and
repaired, involving specialized features of membrane traf-
cking components, in particular the action of dysferlin
upon inux of Ca
21
throughgapsinthecellmembrane
(646,647), as well as other mediators such as the tripartite
motif (TRIM) protein 72 (TRIM72), alternatively called
Mitsugumin 53 (MG53), that also serves as a myokine for
distal organ repair (648). The restoration of membrane
damage is furthermore facilitated by the migration of myo-
nuclei to the site of lesion (649), with this process most
likely supported by the exchange of proteins between
nuclei (650), and the microtubule-dependent transport of
RNAs and ribonuclear proteins within the myober (651).
Then, mechanical strain leads to non-uniformity and over-
stretching of sarcomeres, as well as disruption of Z disks,
resultinginimpairedforceproductionandoverloadofsar-
colemma and T tubules (637). Of note, fast muscle bers
are more susceptible to damage induced by eccentric con-
tractions, which could be due to ultrastructural differences
such as the narrow and more fragile Z disks, suggesting
that a training-induced shift toward slow muscle bers
mightprotectthemusclefromdamage(357). This
raises the question of whether muscles of athletes are
better protected against recurrent mechanical strains
and/or trained muscles have an enhanced repair and
regeneration capacity. Mechanosensing, membrane rup-
ture, stretch-activated channels, and dysregulated ECC
induce intracellular signaling pathways, such as higher in-
tracellular Ca
21
, which result in degradation of damaged
structures as described above. Neurotrophic factors pro-
duced by muscle bers and satellite cells, in particular
activation of the bradykinin receptor B2-nerve growth fac-
tor (NGF) and cyclooxygenase 2 (COX-2)-glial cell line-
derived neurotrophic factor (GDNF) pathways, stimulate
muscle nociceptors and thereby contribute to the pain
experienced in DOMS (637,652). Of note, interference
with the inammatory cascade, even the pro-inammatory
phase, can be detrimental to muscle recovery and adapt-
ive remodeling (637,653). Massage, thermal therapy (hot
or cold), compression, active regeneration, along with vari-
ous pharmacological and nutritional approaches often
have antagonistic effects on DOMS, muscle recovery, and
functional remodeling (654). Thus, probably the best prac-
tice to reduce future/subsequent muscle damage is to
repeat a similar exercise bout, albeit at reduced intensity/
loading (637). Accordingly, in resistance-trained individuals,
recovery of maximum voluntary isometric torque occurs
faster and is accompanied by reduced muscle soreness
and lower CK levels compared with untrained individuals
(655,656). These attenuated symptoms of muscle dam-
age, which also resolve after a shorter period, are hall-
marks of the repeated bout effect(637). A trained muscle
therefore has greater protection against contraction-
induced damage. However, it is not clear whether this
effect is conferred by better resilience against insults,
more efcient repair and regeneration, or a combination of
these (637). The observation of the repeated bout effect
extends to the contralateral, non-trained muscle and sug-
gests a systemic propagation of this signal (637). However,
the underlying mechanisms are unknown. Nevertheless, it
is clear that improved functional recovery is essential to
sustain the high training intensities and volumes of athletes
without overreaching/overtraining (657).
4.8.3. Muscle refueling.
Full functional retrieval requires additional processes,
including restoration of organelle function, sarcomere
repair, and replenishment of substrate stores (FIGURE
17B). Optimally, a supercompensation is achieved, as
observed for muscle glycogen stores (428). The ef-
ciency of refueling depends on promoting an optimal
anabolic environment and is dependent on providing
adequate nutritional availability (81). In this setting,
energy sensors such as AMPK remain inactivated, and
theactivityofanabolicregulatorssuchasmTORare
increased by the availability of amino acids, glucose,
and fatty acids and stimulation by insulin, IGF-1, and
other anabolic hormones (556,557,658). The ensuing
promotion of protein synthesis is instrumental to support
the restoration and de novo formation of sarcomeric and
other structures. Intracellular sensors of glucose and
fatty acids help to restore glycogen and intramyocellular
lipid stores. For example, PGC-1apositively regulates
the transcriptional program for lipogenesis, lipid droplet
FURRER ET AL.
1742 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
assembly, and perilipins (488,659). As noted above,
the increased intramyocellular lipid level in the mus-
cle of endurance-trained athletes resembles the
accumulation of these in muscle from patients with
type 2 diabetes (the athletesparadox). However,
although the diabetic patient is insulin resistant, the
endurance-trained athlete is insulin sensitive based
on the daily turnover and ux of lipid stores (660).
PGC-1aalso increases muscle glucose uptake while
restricting the entry of glucose into glycolysis and
boosting glucose-6-phosphate activity in the pentose
phosphate pathway and glycogen resynthesis (488).
PGC-1amatches these energy-demanding anabolic
processes to adequate mitochondrial function and
ATP synthesis. Moreover, PGC-1a-mediated pentose
phosphate pathway activity produces ribose-5-phos-
phate, the building block for ATP and the other nucle-
otides (488). Intriguingly, the anabolic function of
PGC-1ain this context is opposite to that during mus-
cle contraction, in which this coactivator strongly
stimulates catabolic pathways, including fatty acid
b-oxidation (FIGURE 17C). Thus, to avoid futile cycles
atemporalspecication of PGC-1a, including distinct
transcription factor binding and DNA element recruit-
ment, is likely, although the molecular characteristics
associated with such a specication are unknown.
4.9. Circadian Clock
Almost all physiological processes in humans are under
the control of circadian rhythms (661). Locomotion and
physical activity belong to the most fundamental aspects
of the behavior of higher animals and in an evolutionary
context required coordination of activity with the avail-
ability of prey and food, the avoidance of predators,
along with adequate rest, sleep, and related recovery
processes (662). As described in sect. 2, external cues,
such as timing of food and physical activity, serve as
zeitgebers (time givers) to modulate the circadian clock
in cell-autonomous peripheral tissues and thereby adapt
many tissues/organs to the prevailing environmental
conditions (663). Circadian control of the muscle pheno-
type is inuenced by circadian rhythmicity and executed
by a subset of the transcriptome oscillating in this tissue
(662), with a reciprocal relationship between the core
molecular clock and external stimuli such as the time of
exercise training (FIGURE 18A). Several regulatory
nodes induced by an acute exercise response poten-
tially inuence muscle clock oscillations (FIGURE 18B)
(664666). For example, activation of AMPK affects the
stability of Period2 (PER2) and cryptochrome circadian
regulator 1 (CRY1), two components of the core clock
(662,667). CRY1 and CRY2 interact with PPARb/dand
thereby reduce the oxidative phenotype of muscle cells
(662,667). Circadian Locomotor Output Cycles Kaput
(CLOCK) and Brain and Muscle ARNT-Like 1 (BMAL1),
two other core clock components, regulate SIRT1 tran-
scription in skeletal muscle. SIRT1 in turn deacetylates
PER2 and BMAL1, thus in part counteracting the acetyl-
transferase activity of CLOCK (667). Bidirectional interac-
tions between the clock, mTOR, and protein synthesis
contribute to a link between circadian oscillations and
proteostasis (667). HIF-1ainduces the transcription of
PER1 and PER2, whereas NRF2/NRE2L2 and NF-κB
reciprocally control gene expression of nuclear receptor
subfamily 1 group D member 1 (NR1D1/REV-ERBa)(667).
SIRT1-deacetylated PGC-1ainduces the expression of
retinoic acid-related orphan receptor a(RORa/NR1F1)
and coactivates RORain the transcriptional regulation of
BMAL1 (667). REV-ERBaand RORaare part of an acces-
sory arm of clock control and exert opposite effects in
core clock modulation. CREB activity is affected by CRY1
and in turn controls gene expression of PER1 and PER2
(667). Finally, PPARaand BMAL1 exert mutual transcrip-
tional regulation (662). Many of these interactions have
not been validated in human muscle skeletal, but there
is ample potential for cross talk between factors that are
modulated by contractile activity and those that entrain
and synchronize the molecular clock (662,663,665
667). Accordingly, the gene expression of several com-
ponents of the molecular clock are affected by exercise,
such as BMAL1, PER2, and CRY1 (663). Curiously, other
clock components, including CLOCK, PER1, CRY2, and
REV-ERBb/NR1D2, were unaffected by skeletal muscle
contractile activity in these studies, which conicts with
a general modulation of clock phase and amplitude by
exercise. As such, these data suggest a specic control
of a subset of clock genes and not a complete resynch-
ronization. Such a targeted interaction could imply that
some of these proteins might have additional functions
unrelated to the molecular clock such as the control of
cellular metabolism. Indeed, the consequence of the
cross talk between exercise factors and the molecular
clock for circadian rhythmicity, muscle function, and
exercise adaptation is unclear and is likely to be con-
founded by many factors including meal time, sleep,
psychological stress, and other zeitgebers (665). Animal
experiments with voluntary wheel running conducted
under a skeleton photoperiod suggest that skeletal mus-
cle clock oscillations are robust even in the face of per-
turbations induced by daytime running and feeding
(668). Moreover, studies of the transcriptome, proteome,
and phosphoproteome of skeletal muscle of mice
undergoing maximal endurance tests at different times
of the day indicate a strong association with metabolic
parameters such as muscle and liver glycogen concen-
trations, suggesting more indirect effects of circadian
oscillations (668). Thus, future studies need to determine
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1743
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
the extent of the inuence of the molecular clock on the
exercise responses and vice versa, including functional
consequences for acute endurance and resistance exer-
cise bouts, chronic training adaptations, and peak per-
formance at competitions.
4.10. An Integrative View of the Molecular
Mechanisms
In this section, we summarized the molecular mecha-
nisms and pathways that are activated in skeletal mus-
cle in response to endurance and/or resistance
exercise, with a focus on the putative pathways for
whichanassociationbetweentheacuteexercise
response and chronic training adaptation has been
consistently reported. Other reviews have discussed
the exercise-induced changes in various pathological
conditions such as muscle wasting, cachexia, or sar-
copenia (669672), which may be completely distinct
or share some commonalities with those in healthy
muscle in well-trained athletes. There are also several
pathways and factors that contribute to the pathoetiology
of muscle wasting and diseases. Although voluntary and
forced muscle inactivity do not necessarily mirror con-
tractile engagement (673), some of these could play a
role in exercise training adaptation. For example, the lev-
els of myostatin, which rise in several pathological states
(15), are reduced by exercise (674), and the absence of
myostatin signaling may be an important contributor to
training adaptation. However, this might depend on fac-
tors such as baseline control, exercise modality, or the
pathological context regarding comorbidities. Related
mechanisms such as the repression of activin receptor
type II (ActRII), for which myostatin is one of the ligands,
by the m(6)A methyltransferase-like-3 (METTL3) could
also potentially contribute to muscle hypertrophy after
exercise training (675). Readers are referred to recent
reviews on muscle atrophy and disease states (560,
676679). Finally, many studies using gain- and loss-of-
function of targeted genes have reported altered muscle
metabolism or contractile function, most of which require
validation after exercise interventions in both rodent mod-
els and humans (485,486,680,681). Many of these fac-
tors will undoubtedly add to our current knowledge
regarding human exercise-induced muscle plasticity.
Notwithstanding these limitations and caveats, we still
have a poor understanding of how exercise adaptation is
regulated. Most studies focus on individual pathways and
factors, often centered on the usual suspects,and pre-
cisely how the complex and interdependent network of
signaling pathways and mechanisms is spatio-temporally
coordinated and integrated is unclear (FIGURE 19). A
reductionist study of these pathways is difcult because
muscle activity and plasticity are under robust control,
with numerous redundancies (overlapping or parallel
pathways that are engaged in a physiological setting),
backups/contingencies (pathways that are engaged if
other mechanisms are impaired, controlling the same
A
B
Central clock
Metabolism
Contraction
CLOCK
BMAL1
E boxE box
PER
CRY
ROR
REV-ERB
CRY PER
RORE
BMAL1
ROR
AMPK
PPARβ/δ
SIRT1
HIF-1α
mTOR
PGC-1α
CREB
PPARα
Activity levels
and food intake
Skeletal muscle
core clock
NFE2L2
NF-NB
REV-ERB
FIGURE 18. Circadian regulation of skeletal muscle function and
plasticity. A: circadian regulation of activity and muscle function.
Circadian regulation of sleep-wake cycles, feeding, physical activity,
or other zeitgebers is sensed and translated by the master clock in
the suprachiasmatic nucleus as well as peripheral clocks in almost
every cell in the human body. As a consequence, various physiologi-
cal processes, e.g., muscle cell metabolism or contractile perform-
ance, can be affected in a circadian manner. B: interactions between
core clock genes/proteins with regulators of muscle function and
plasticity. Such cross talk pertains to the regulation of gene expres-
sion, protein-protein interactions, and/or enzymatic activity. Thereby,
physiological processes in muscle might be affected by the core
clock. Inversely, muscle ber metabolism, contractile activity, oxygen
availability, redox balance, and other perturbations could modulate
the skeletal muscle clock. AMPK, AMP-dependent protein kinase;
BMAL1, brain and muscle ARNT-like 1; CLOCK, circadian locomotor
output cycles kaput; CREB, cAMP response element binding protein;
CRY, cryptochrome circadian regulator; HIF-1a, hypoxia-inducible
factor 1a; mTOR, mammalian target of rapamycin; NFE2L2, nuclear
factor erythroid-derived 2-like 2; NF-κB, nuclear factor κB; PER, pe-
riod; PGC-1a, peroxisome proliferator-activated receptor ccoactiva-
tor 1a; PPARa/b/d, peroxisome proliferator-activated receptor a/b/d;
REV-ERB, nuclear receptor subfamily 1 group D member 1/2; ROR,
retinoic acid-related orphan receptor; RORE, ROR elements; SIRT1,
sirtuin 1. Image created with BioRender.com, with permission.
FURRER ET AL.
1744 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
biological program), and alternatives (processes that are
used when others are dysfunctional, leading to different
physiological adaptations but resulting in similar out-
comes) (18,41,682). For example, three of the most char-
acterized factors with putative roles in contraction-
inducedremodeling(AMPK,mTOR,andPGC-1a)aredis-
pensable for many training-induced adaptations. First, in-
ducible ablation of muscle AMPKadoes not affect whole
body substrate utilization, muscle glucose uptake, fatty
acid, or mitochondrial respiration during exercise (683).
Similarly, muscle-specicknockoutmiceforPGC-1astill
increase mitochondrial biogenesis with training, at least
in some studies (576). The function of mTORC1, with
roles in the early events leading to muscle hypertrophy, is
compensated by non-mTOR-dependent pathways during
recovery from exercise (581). Integrity of raptor is impor-
tant for hypertrophy induced by synergist ablation over-
load but not the related boost in the rate of protein
synthesis (684). Genetic models with inducible muscle-
specic inhibition by ablation of raptor and sustained acti-
vation of mTORC1 by disruption of TSC1 have little effect
on preserving muscle mass (524) or facilitating hypertro-
phy (685), respectively. It is conceivable that such results
are caused by awed or imperfect experimental model
systems, with many studies relying on constitutive gain-
and loss-of-function. Moreover, targeting strategies might
be imprecise, such as the use of raptor and rictor to ge-
netically ablate mTORC1 and mTORC2, respectively, lev-
eraging TSC1/2 to activate mTORC1 activity, or using
rapamycin as a pharmacological inhibitor of mTORC1.
CnA CaMKII
PGC-1α
AMPK mTOR
SIRT1
cAMP
β2-AR IGF-1, insulin
ROS
NO
ATP/ADP/AMP
Metabolism Vascularization
Mechanosensing
?
Spatio-temporal signal coordination and integration?
Adipo-, hepato-, osteokines
e.g., adiponectin
Ca2+
Stress kinases
e.g., p38 MAPK
NADH/NAD+
CARP
DARP
ANKRD2
branched chain
amino acids
CREB, ATF2
NFAT
MEF2 HDAC5
NCOR1
TFEB
MEF-2
MondoA
HSF1
HIF-1α, HIF-2α
FOXO3A
ATF4, ATF5, ATF6, CHOP
PER, CRY1, BMAL1, CLOCK, REV-ERB
NR4A2 (NURR-1), NR4A3 (NOR-1)
PPARβ/δ, PPARα
ERRα, ERRγ
NRF1, GABP, TFAM
GR
SRF, AP-1, EGR-1
TAZ/YAP/TEAD1
NRF2/NFE2L2
Contractility
Fiber type
Motor unit
remodeling Myokines
Myobolites
Proteostasis
Autophagy
Redundancies
Contingencies
Alternative pathways
Output specification and coordination?
TF
FIGURE 19. Molecular mechanisms in contracting muscle bers. A small selection of molecular sensors and mediators as well as a simplied sum-
mary of the presumed interactions are depicted. Nota bene: the spatio-temporal integration and coordination of these pathways, functional redundan-
cies (e.g., overlapping or parallel pathways), contingencies (back-up processes with the same function), alternatives (back-up processes leadingto
similar adaptations using different functions), specication and coordination of downstream effects and adaptations, in particular in chronic settings, as
well as many other aspects are still only understood at a very rudimentary level. b-AR, b-adrenoreceptor; CARP, cardiac ankyrin-repeat protein; CnA,
calcineurin A; CREB, cAMP-dependent binding protein; DARP, diabetes-related ankyrin-repeat protein; ERRa/c,estrogen-relatedreceptora/c; GR, glu-
cocorticoid receptor; HSF1, heat shock factor 1; NO, nitric oxide; PPARa/b/d, peroxisome proliferator-activated receptor a/b/d; ROS, reactive oxygen
species; SIRT1, sirtuin 1. See text for other abbreviations. Image created with BioRender.com, with permission.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1745
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
Even though raptor and rictor are unequivocal members
of mTORC1 and mTORC2, many other potential interac-
tions of these proteins beyond their direct function in
forming the mTOR complexes have been proposed (686)
or demonstrated (687). Similarly, TSC1/2 and the down-
stream effector RHEB clearly result in activation of mTOR
but potentially also the modulation of the function of other
proteins (688). Furthermore, physiological mTOR activity
is regulated transiently and in a pulsatile manner, quite dif-
ferent from the sustained, long-term loss- and gain-of-
function in the animal models. Finally, rapamycin is a
potent inhibitor of mTORC1 activity but at different dos-
ages and administration durations serves also as a modu-
lator of the activity of other protein complexes, including
mTORC2 and signal transducer and activator of transcrip-
tion 3 (STAT3) (689). Notably, STAT3, like mTORC1, exerts
effectsonmusclemass(690). However, these seemingly
paradoxical ndings in relation to PGC-1a,AMPK,and
mTOR should not necessarily be interpreted as a sign of
their dispensability or insignicance in the physiological
exercise response but rather as proof of the resilience of
whole body systems to adapt even under adverse condi-
tions. For example, although muscle-specicPGC-1aknock-
out animals can partially adapt to exercise training, this
adaptation differs from the physiological training response
seen in wild-type animals and occurs despite a lack of vas-
cularization or metabolic adaptations in lactate and ketone
body metabolism, thus presumably relying on alternative
processesthatprovidesimilarbenets (575577). How
such compensation is achieved is unclear, but it is conceiva-
ble that the transcription factor binding partners still regulate
the corresponding target genes even in the absence of this
and other coregulators, albeit at lower levels or altered tar-
get specicity (474,486). PGC-1acan regulate target genes
with different binding partners, such as those involved in
the hypoxic response by coactivating AP-1 or ERRa,imply-
ing functional redundancies in the transcriptional network
engaged in exercise adaptation at different levels (474).
Future studies should aim at obtaining an unbiased, holistic
picture of the molecular mechanisms that control muscle
plasticity in response to endurance and resistance training
(18,40). Such insights might also be important to better
understand training interference effects and help in the
design of concurrent training programs for athletes involved
in multisport events (i.e., the triathlon) or who require traits of
both endurance and power for successful performance.
Finally, our knowledge about the mechanisms that control
chronic adaptations to endurance- or resistance-based
training is sparse compared with that describing the
responses to a single bout of exercise. Many of the adapta-
tions in elite athletes with a prolonged history of training are
not reected in the transcriptional changes that occur after
one acute exercise bout, such as the change in expression
of different myosin heavy chains (9). Therefore, it is unlikely
that a trained muscle only reects the additive or sustained
changes that are induced in individual acute exercise bouts
(9). Epigenetic changes might be involved in priming and/or
altering the gene expression prole of trained versus sed-
entary and of acute exercise-induced changes in the trained
versus the naive state (41,585,691). Precisely how the tem-
poral sequence of individual exercise bouts over time
results in training adaptation remains a fertile area for future
research. Together, sects. 24 have described training
practices and the physiological and molecular pathways
involved in the acute exercise response; sect. 5 integrates
these aspects to discuss interindividual differences and the
specic aspects that underpin elite athlete performance.
5. CAN WE ALL BECOME GOLD MEDALISTS?
5.1. Individual Responses to Training
Elite sporting performance is the result of the interaction
between genetic and training-related factors (692,693),
and although several genes or gene clusters have mod-
erate associations with performance or performance-
related phenotypes (694696), the genomic signatures
associated with elite athletic performance across a wide
range of events/sports have yet to be identied.
Notwithstanding the lack of direct links between genetic
variants and elite performance, the notion that there exist
interindividual responses to exercise training and that
innate factors may explain a large part of the training-
induced variance in maximal aerobic capacity in previ-
ously untrained persons can be traced back to the classic
HERITAGE studies conducted in the 1980s by Bouchard
and colleagues (697). In these and other recent investiga-
tions, the most common reported primary outcome mea-
sure in response to endurance-based training programs
was an individuals_
VO2max attained during an incremental
exercise test to exhaustion, either treadmill walking/run-
ning or ergometer cycling, usually lasting 1015 min.
After 23 mo of endurance training (3 or 4 sessions
per week), _
VO2max is typically increased by 1015% at
the group level, but the magnitude of improvement
canbeaslittleas12% and as great as 35% (698). A
key observation from these early human training stud-
ies was that up to 20% of subjects demonstrated little
change in _
VO2peak in response to a standardized train-
ing protocol and were considered exercise resistant
(699). However, individuals who demonstrate a low
training response to one outcome measure (e.g., _
VO2max)
do not always display the same response in other
FURRER ET AL.
1746 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
parameters. This makes the concept of responders versus
non-responders open to different interpretations. Indeed,
exercise training induces a multitude of health- and per-
formance-related benets (9,700), some of which may
even have a non-physiological basis (701).
Recently, the notion that there exist exercise-resistant
subjects or non-responders has been challenged on the
grounds that in those studies in which individuals exhibit
no meaningful change in a specied outcome variable,
the training impulse has been inadequate in terms of ei-
ther volume or intensity overload (702). To test this hy-
pothesis, Montero and Lundby (702) recruited 78 healthy
young male volunteers and rst subjected them to 6 wk of
supervised training. Individuals trained in ve groups dif-
fering in the number of exercise sessions per week: they
performed 60 min of cycle ergometer exercise either one,
two, three, four, or ve times a week, corresponding to
60, 120, 180, 240, and 300 min total training time. After
completing this rst phase of training, the group mean
maximal work rate sustained during an incremental cycle
test (W
max
) was increased in all groups apart from the
group who undertook just a single training session per
week. In groups one, two, three, four, and ve, 69%, 40%,
29%, 0%, and 0% of individuals, respectively, were non-
responders [i.e., did not have a signicant increase in the
maximal work rate (W
max
)attainedduringanincremental
test to exhaustion]. Subsequently, individuals classied as
non-responders to the rst phase of training performed a
second 6-wk block of training in which two additional 60-
min sessions per week were included, independent of the
number of sessions completed during the rst 6 wk.
Following this protocol of volume overload, the lack of
training-induced increase in W
max
was universally abol-
ished. Although further research needs to be undertaken
to determine whether such ndings can be extrapolated
to other populations, these results fundamentally chal-
lenge the notion of non-responders and suggest that if a
training stimulus is of sufcient volume and/or intensity
and follows the principles of progressive overload and
specicity, then those individuals thought to be exercise
resistant or low responders can indeed become res-
ponders(702). In this context, Booth and Laye (703)
believe that the term non-responder should be replaced
by individuals who demonstrate low sensitivityto a
training stimulus and that such individuals merely require
increased training volumes and/or intensity to drive
favorable responses. At the population level, focusing on
only selected measures of training response and label-
ing an individual as a non-responder is somewhat of a
reductionist approach to exercise. Indeed, if we accept
that exercise is a polypill(46,704) exerting a plethora
of positive benets (700), then by focusing on a small
number of measures of response we ignore the fact that
exercise works through so many different pathways and
mechanisms that the chances of an individual exhibiting
no single biological benet is highly unlikely (701,705).
Regarding elite athletic performance, the issue of low
responders or non-responders becomes somewhat irrel-
evant: to become an elite performer, one must be initially
well endowed in the traits that have critical roles in the
athletes event, and one should also be a high responder
to exercise training (83). Furthermore, without a rich
genetic endowment, world-class athletic performance is
unattainable (83). In this regard, there are substantial dif-
ferences in performance-related traits measured in the
sedentary state (i.e., before any training intervention).
Individuals who have high levels of a trait before expo-
sure to exercise training are greatly predisposed to expe-
riencing early success, which also might have a big
impact on motivation and subsequent training adherence.
This is not the case in non-elite sedentary populations in
which no correlation is observed between the baseline
level of tness (i.e., _
VO2max) and the response to an exer-
cise training dose (699), suggesting a unique biology
underlying trainability. In support of this contention,
Rønnestad et al. (292) report the case of an individual
with a _
VO2max of >70 mL/kg/min in the untrained state
that increased by 30% after 3 yr of specialized training to
>95 mL/kg/min and coincided with a world championship
title. Animal studies of selected breeding for aerobic run-
ning capacity also reveal high and low responders to
standardized exercise training. To determine the inher-
ited components of acquired running capacity, a model
of two-way articial selection for animals that were either
low or high responders to exercise training was devel-
oped (706). These two animal lines were tested for maxi-
mal treadmill running distance before and then after 8 wk
of standardized treadmill run training protocol (i.e., the
same absolute training load). After 15 generations of
selection, and 8 wk of training, the high-response rats
improved an average of 223 m in run distance, whereas
exercise capacity declined by 65 m in the low-response
animals. Taken together, a continuum of responses to
standardized exercise training protocols exists (698), yet
low sensitivity to adapt may be mitigated by revision of
exercise prescription including training frequency, dura-
tion, and intensity (701,702). Although an understanding
of how genetic factors contribute to interindividual train-
ing responses may improve personalization of training
prescription, at present genomic predictors for response
trainability are lacking (707).
5.2. Genetic Predisposition to Becoming an Elite
Athlete
An indication of genetic predisposition toward athletic
prowess could be inferred from a variety of observa-
tions. For example, people from the western parts of
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1747
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
Africa (including Ghana and Nigeria) as well as the de-
scendants of the slaves that were transported from
these regions to the New World(the West Indies or
the United States) are excellent sprinters. In contrast,
athletes from Eastern Africa, such as Kenya and
Ethiopia, are famous for their extraordinary long-dis-
tance running feats. In fact, 90% of elite marathon
runners worldwide are of Ethiopian or Kenyan descent
(708). However, it is unclear whether different genetic,
and in extension anthropometric (709)orratherenvi-
ronmental, factors and training practice (118)explain
this phenomenon. For example, the higher altitude
plateaus of the Eastern African countries might facili-
tate endurance training adaptations in contrast to the
mostly sea-level landscape of Western Africa. Despite
these alternative explanations, it is estimated that
65% of athletic ability can be explained by genetic
factors (710). Moreover, there are data indicating that
maximal endurance capacity as well as trainability is
inherited (699,711) and that the genetic makeup accounts
for a substantial contribution to performance levels (712).
However, despite evidence that genetic components are
strongly related to the phenotypic traits of elite athletes,
knowledge of the specic genes underpinning this pre-
disposition is limited. Rare examples for extreme genetic
variants underline the genetic contribution to athlete sta-
tus. One example is the mutation in the EPO receptor
gene that resulted in a more active truncated protein in
the Finnish cross-country skier Eero M
antyranta (713). His
Hb levels were at least 200 g/L, which is substantially
higher compared with other endurance athletes or non-
athletes (150 g/L) and could thereby have contributed
to the three Olympic gold medals and two World
Championships he won over his career (266,713). In
broader populations, the two polymorphisms that are
most described and linked to athletic performance are
located within the ACE and ACTN3 genes that encode
for the angiotensin I-converting enzyme and a-actinin-3,
respectively (714). In fact, the rst polymorphism that was
associated with athlete status was ACE I/D (715). ACE is
part of the renin-angiotensin system and is involved in
regulating blood pressure by converting angiotensin I to
the active vasoconstrictor angiotensin II. ACE activity in
serum is lower in the presence of the insertion (I) allele
containing 287 base pairs within intron 16 (716). The I al-
lele is associated with successful endurance capacity,
whereas the deletion (D) allele is associated with prowess
in strength/power events (716). However, inconsistencies
exist within this classication: in elite endurance runners,
an association with the D allele or the I allele as well as a
lack of any association have been reported (708,714). In
comparison, data on ACTN3 polymorphisms are more ro-
bust. ACTN3 is an actin-binding protein that is exclusively
expressed in type II bers and located at the Z disk,
suggesting a role in high-velocity force contractions (717).
The single-nucleotide polymorphism (SNP) in the ACTN3
gene results in a premature stop codon (X) instead of the
arginine (R) at position 577, and the XX genotype is de-
cient in expressing the ACTN3 protein (718). The 577R al-
lele has been associated with elite sprint/power athletes
and explosive performances, RR being superior to RX
and XX genotypes (719,720). In contrast, the XX variant is
more frequently observed in elite endurance athletes
compared with non-athletes and is extremely rare in elite
power athletes (719). In line with these observations, the
absence of ACTN3 results in an increased endurance
performance and higher oxidative phenotype in the mus-
cle of knockout mice (721).
During the last two decades, at least 155 polymor-
phisms related to elite endurance (93 polymorphisms) or
power athletes (62 polymorphisms) have been identied
(722). Endurance markers that have been replicated in at
least three independent studies include ACE I, ACTN3
577X, HFE (homeostatic iron regulator), PPARA, and
PPARGC1A, and power markers include ACE D, ACTN3
577R, AMPD1 (adenosine monophosphate deaminase 1),
HIF1A, MTHFR (methylenetetrahydrofolate reductase),
NOS3, and PPARG (722). Of all 93 polymorphisms associ-
ated with elite endurance athletes, the 3 located within
the ADRB2 (adrenoceptor b2),ADRB3,orPPARGC1A
genes have also been shown to be associated with base-
line _
VO2max of a non-athlete population (723)andthe5
variants of the ACE, AMPD1, CKM, HIF1A, and PPARD
genes with _
VO2max trainability (724). However, even
though several genetic variants have been identied
in genomewide association studies (GWASs) of elite
athlete status, there is no subset of genes that allows the
identication of elite athletes (696). Often, studies are
underpowered, which is not surprising regarding the lim-
ited number of elite athletes and accessibility of biological
samples, and therefore many results cannot be replicated
in different cohorts (696). Based on a meta-analysis
including 1,520 endurance athletes and 2,760 non-ath-
letes control subjects, a polymorphism in the GALNTL6
gene, encoding for polypeptide N-acetylgalactosaminyl-
transferase-like 6, may be a signicant marker for athletic
performance (696). In endurance athletes, the C allele is
overrepresented compared with non-athletes, whereas
the T allele of GALNTL6 is more frequently observed in
elite power athletes compared with endurance athletes
or non-athletes and is associated with higher peak power
in active men (696,725). Another recent meta-analysis in
elite endurance athletes identied polymorphisms in the
MYBPC3 (myosin-binding protein C3) and NR1H3 (nuclear
receptor subfamily 1 group H member 3/LXRa) genes that
were also correlated with _
VO2max (726). SpecicSNPsin
the HFE, NFIA-AS2 (nuclear factor I A antisense RNA 2),
and TSHR (thyroid stimulating hormone receptor) genes
FURRER ET AL.
1748 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
that are more frequently observed in elite endurance ath-
letes compared with controls have also been associated
with high _
VO2max among athletes (727729). Additionally,
athletes with homozygous C alleles of NFIA-AS2, encod-
ing a long noncoding RNA, have improved hematologic
parameters such as higher Hb levels, since NFIA-AS2
may be involved in the regulation of the transcription fac-
tor NFIA and erythropoiesis (727,728). However, even if
a combination of GWAS and selected physiological
measures in elite athletes may help identify SNPs for vari-
ous genes, data including such measures are limited and
so far these candidates have not been replicated (722). In
the future, it might be possible to link polymorphisms to
the responsiveness of individuals (i.e., low vs. high res-
ponders) to standardized training interventions, as a set of
21 SNPs was able to predict almost 50% of the variation
observed in the _
VO2max training response (698). However,
as noted, replication of these data remains difcult, and
not all SNPs seem to affect trainability across all popula-
tion groups (698). Importantly, the identied polymor-
phisms are based on associations, and for the large
majority of these the functional relevance and mechanistic
aspects of the gene variants in muscle biology are
unknown. Therefore, identication of SNPs with larger
effect sizes, replication of the identied SNPs in independ-
ent cohorts, as well as studies including a larger sample
size and possibly additional physiological measures to dis-
criminate individuals (i.e., _
VO2max and performance out-
comes) are necessary to gain knowledge about genetic
factors that underpin elite athlete performance. These
studies will have to be combined with mechanistic investi-
gations of the functional effect of gene variants and SNPs
to understand how differences are brought about.
Gain- and loss-of-function studies in mouse models
have identied 31 genes associated with endurance
performance (681). For eight of these, genetic var-
iants have been reported to be associated with elite
endurance athletes (681,722). The genes associated
with an elite endurance athlete status and enhanced
endurance performance in a gain-of-function mouse
model include PPARD, PPARGC1A, PPARGC1B, and
PPP3CA (protein phosphatase 3 catalytic subunit a),
and those correlated with endurance performance in
a loss-of-function mouse model include ACTN3,
ADRB2, BDKRB2 (bradykinin receptor B2), and HIF1A
(681,722). Although 47 genes were identied in mouse
models that induce hypertrophy in a gain- or loss-of-func-
tion model, no corresponding human polymorphisms
were described for most of the genes that were associ-
ated with a strength/power muscle phenotype (680).
However, gene variants in IGF1 and ADRB2 have been
found in the phenotype of power athletes and are 2 of
the 47 genes that induce hypertrophy in a gain-of-func-
tion model (722).
Collectively, most of the identied polymorphisms dif-
fer between studies and only explain a very small frac-
tion of the interindividual differences in endurance and
strength. As a consequence, the current knowledge is
inadequate for talent identication or prediction of train-
ing response (68). Besides the question of whether
genetic talent identication will ever be feasible in the
future, ethical and practical issues also need to be con-
sidered (718,730). Geneticprediction of athletic prow-
ess and specialization disregards personal preferences
and choices, with potential detrimental consequences
on enjoyment, motivation, and ultimately adherence.
Talent identication and premature specialization might
also preclude the multidisciplinary practice in youths
that predicts world-class performance (731). Moreover,
sensitive genetic information has potential for misuse
and unexpected outcomes, and could have psychologi-
cal consequences that could even extend to other family
members. Additionally, the reported associations are
observed at the population level and hence have a very
low predictive value at the individual level (732). In fact,
there are frequently individuals with a seemingly less
favorable genotype who achieve elite athlete status
(718). Moreover, monozygotic twin studies revealed a
strong impact of discordant leisure-time physical activity
on performance, tness, health, and well-being, to a
large extent disconnected from the genetic endowment
(733). In summary, the current scientic evidence sup-
porting the contributions of specic genetic variants to
elite athletic performance phenotypes is weak (734).
This is partly because complex traits are modulated not
by several genes with large effect sizes but instead by
polygenic systems dened by hundreds or thousands of
loci, characterized by alleles with small effect sizes, plus
less frequent alleles (83). In contrast to the commercial
for-prot genetic testing to predict training selection and
response, non-genetic testing, such as assessment of
physical performance, might be more useful for athlete
stratication (68,83,692,693,735). Importantly, instead
of talent identication, genetic information could also be
used for the screening of polymorphisms that are associ-
ated with injury risk among athletes (i.e., stress fractures
or tendinopathies) to reduce injury by individualized pre-
ventative measures (718), with much fewer ethical consid-
erations attached.
5.3. The Aging Athlete and Athletic Performance:
Slowing Down with Speed
During the past century, there has been a steady
increase in life expectancy among most countries in the
Western world. However, such enhanced longevity (life
span) has not always been accompanied by a propor-
tional elevation in healthy life expectancy, so-called
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1749
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
healthy agingor healthspan(736). Thus, an objective
of medical and aging research is not necessarily to pro-
long life span per se but instead to increase the health-
span and compress morbidity later in life (737). At the
population level, aging is strongly associated with a rise
in sedentary behavior and concomitant declines in phys-
ical and mental capacity. An examination of the perform-
ance proles of individuals who continue to train and
compete throughout their entire life provides insights
into the extremes of human function and the upper limits
of physiology during the human aging process (738,
739). The birth of the mastersathlete movement (those
individuals >3540 yr of age, depending on the sport)
can be traced back to the late 1960s and early 1970s. At
this time, there was a massive increase in the number of
people who started exercising, either for health and
pleasure or to pursue competitive endeavors. In the
United States, this escalation in structured physical activ-
ity initially centered on distance running and was
inspired by a few select individuals such as Roberta
Gibb and Kathrine Switzer, who were the rst female n-
ishers in the Boston marathon in 1966 and 1967, respec-
tively, and Frank Shorter, who won the gold medal in the
mens Olympic marathon in 1972 and silver in the 1976
Olympics. Throughout the next two decades, there was
an explosion in the number of marathon races held in
capital cities throughout the world, with applications to
run in these events far exceeding the number of avail-
able starting places. The mid-1970s also saw the birth of
the triathlon, which consisted of an amalgamation of
three separate sports: swimming, cycling, and running.
From humble beginnings (15 men started and 12 nished
the inaugural Ironmantriathlon in Hawaii in 1978), the
Hawaii Ironman is generally considered one of the most
difcult 1-day sporting events in the world, and today
Ironman races attract almost half a million entries world-
wide each year. These competitive, mass-participation
inner-city marathons and triathlons were the motivation
for a generation of women and men to start training for
specic competitions/races and laid the foundation for a
generation of athletes who are now in their eighth or
ninth decade of life and have participated in formal exer-
cise training throughout their life span (740). As such, we
now have both cross-sectional and longitudinal data on a
cohort of masters athletes who have maintained rigorous
training schedules over many decades and have better
health outcomes than their age-matched non-athletic
counterparts. Such well-trained individuals represent the
optimum phenotype for examining the effects of aging on
performance and vice versa, as these individuals are
most likely minimally affected by the negative effects of
the age-related decrease in voluntary physical inactivity
and changes thus largely driven by an inherent aging pro-
cess (739). Although cross-sectional data on masters
athletes are easier to obtain than longitudinal data, the
former can only provide the age-related performance
decline for a population, whereas longitudinal data show
individual trajectories. A detailed analysis of the age-
related declines in performance across multiple sporting
events is beyond the scope of this review, and the reader
is referred to previous work in this area (741746). Here,
we provide a general overview of the effects of aging on
overall physical performance declines and discuss some
of the mechanisms that underpin this decay in perform-
ance capacity.
Plotting age-group world records for males and
females across various sporting events provides insights
into the rates of decline in performance capacity across
the healthspan (FIGURE 20). Such cross-sectional data
offer the best-case scenariofor each age band but do
not provide any information on the individual athletes
decay curves. Several observations are noteworthy.
First, the rate of performance decline does not appear
to differ between or within sports (e.g., the various track
and eld events and multiple swimming strokes and dis-
tances) (745). Second, there are no major sex-related dif-
ferences in the deterioration in performance for most
sports (746), and although there is a greater absolute
drop-off in performance for females compared to males,
the relative decline is similar. Performance declines for
most of the running, swimming, and cycling events, inde-
pendent of distance, are not linear but curvilinear (741).
However, at 70 yr of age, there is an accelerated
increase in performance decrements for almost all sport-
ing performances. For example, world records show
rapid declines after age 70 in swimming, long-distance
running, and sprint performance (744). Although it is
unlikely that those athletes still training and competing
after the age of 70 can maintain the same absolute train-
ing volumes and intensities, there is no reason to sus-
pect that the relative training intensity has diminished. It
is worth noting that only 5% of athletes competing at
age 80 yr are still competing at age 90 yr. Therefore,
world-best performances in these latter years may
merely reect a lower number of participants rather than
any underlying physiological factor responsible for the
drop in performance capacity. Nevertheless, the per-
formance decline in athletes aged 80 yr is threefold
greater compared with athletes aged 3069 yr (1.62%
vs. 0.46% per year) and accelerates around 67 yr, espe-
cially for sprint/power disciplines (742,743).
The factors that constrain performance with aging are
event-specic, meaning that for each sport/event there
may typically be one or more physiological/mechanical
systems that limit exercise capacity. For example, the
decline in maximal running velocity (independent of dis-
tance) is likely to be underpinned by deteriorations in sev-
eral properties in skeletal muscle that include a decrease
FURRER ET AL.
1750 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
in the maximum strength or power output, a slower rate of
force development and transmission, and a reduction in
elastic energy storage and recovery of tendons (748).
During the normal aging process, there is a progressive
loss of muscle mass, mainly in the lower body, which
increases after the age of 45 yr, with the absolute
decline being faster in men than in women (749). Even in
masters athletes who have undertaken a lifetime of train-
ing, a reduction in size and function of muscle is observed,
which is most pronounced in the fast bers (750). This
selective loss of fast-twitch bers explains, in part, the
greater magnitude of performance decline in sprint and
explosive events compared with endurance-based activ-
ities. Cross-sectional data from 37- to 90-yr-old masters
athletes indicate that peak anaerobic power declines by
714% per decade, with this reduction being similar
between male and female athletes (751). Despite the
selective loss of fast muscle bers, Cristea et al. (752)have
shown that a 20-wk program that combined sprint training
with heavy and explosive strength exercises specically
targeting fast-twitch muscle bers improved maximal, ex-
plosive, and sport-specic force production in elite mas-
ters sprinters (aged 5278 yr). Furthermore, at a single-
ber level, it seems that power output per unit size (i.e.,
muscle quality) is not reduced in muscles of aging
athletes (753), suggesting that the loss in power out-
put can mainly be ascribed to the loss of fast muscle
bers rather than muscle quality. In contrast to
sprint-based activities, men and women with a pro-
longed history (5 days/wk for 7 h/wk over the past 52
yr) of endurance-based training have capillarization
and aerobic enzyme activity similar to younger (25 yr
of age) exercisers, which in turn is 2090% greater
than the parameters determined in elderly age-
matched non-exercisers. Although whole muscle
responses offer unique insights into age-related
muscle deterioration, they fail to provide a better
understanding regarding the potential mechanisms
underlying this phenomenon.
With endurance-based sports, there is a decrease in
_
VO2max of 1%peryearaftertheageof35yr(754). This
is due to a combination of factors including, but not lim-
ited to, an increase in ventricular stiffness that contributes
to worsening left ventricular diastolic function reected
by reductions in early inow velocity, ratio of early to late
inow velocity, and early diastolic tissue velocity and
increases in the isovolumic relaxation time and the time
constant of isovolumic pressure decay (755). Although
lifelongexercisershaveagreaterstrokevolumeandcon-
sequently a superior functional capacity and cardiovascu-
lar reserve than their sedentary peers, there may be
some adverse consequences of a lifetime of rigorous en-
durance-based training. Several studies have reported
increased risk of atrial brillation, atherosclerotic plaques,
and aortic dilation in masters athletes compared with age-
matched untrained individuals (756758). A J-shaped
association between exercise volume and all-cause mor-
tality, cardiovascular disease, and total cancer, as opposed
to the L-shaped association with diabetes, would imply a
negative risk prole at very low and very high volumes of
activities (759). These ndings, however, contrast with
results of other studies in different athlete populations, in
which a J-shaped association was not observed (760,
761). In fact, many studies and meta-analyses reveal
reduced mortality in former athletes (761763), even in
those competing at the highest level such as United
States, French, or Polish Olympians or French participants
oftheTourdeFrance(764767). Often, assessment of
potential effects of genetic components and the benets
of lifelong training in elite athletes is confounded by
healthier lifestyle habits such as being more active or a
non-smoker (768,769).
9
8
7
6
5
4
3
2
1
0
35-39
40-44
45-49
50-54
55-59
60-64
65-69
70-74
75-79
80-84
85-89
90-94
24
22
20
18
16
14
12
10
8
0
35-39
40-44
45-49
50-54
55-59
60-64
65-69
70-74
75-79
80-84
85-89
90-94
Age group Age group
Time (h)
Time (s)
Marathon records 100 m records
Masters athletes
Women
Men
Women
Men
World record World record
FIGURE 20. Age-related decline in records of sprint and endurance events. Records for marathons (https://world-masters-athletics.com/
championships/non-stadia/) and 100 m sprints (747) of masters athletes represent the age-induced reduction in tissue and organ function that despite
high training loads leads to a decrease in performance. However, they also highlight the potential of the human body to achieve incredible performan-
ces at advanced age with adequate training.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1751
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
5.4. Looking Ahead: Personalized Training Using
Wearables
The application of wearable appliances/sensors to sport
is a relatively new phenomenon. In 1978, a Finnish com-
pany pioneered the rst wearable heart rate monitor,
subsequently introducing a monitor with an integrated
computer interface, which gave athletes the ability to
view and analyze their training data on a computer for
the rst time. In 2009, a professional European soccer
club used wearable devices for measuring player work-
load during games. That device was among the rst to
allow coaches real-time monitoring of each players bio-
metrics for signs of exhaustion or injury while on the eld
of play. The generation of real-time data during competi-
tion situations is important: athletes have long been able
to monitor their physiological status under controlled
laboratory conditions, but competition demands includ-
ing prevailing environmental factors (heat and humidity,
wind speed, altitude, other athletes/players, crowds)
impose a different set of stressors that cannot be mim-
icked in the laboratory (770). Over the last 20 years, the
growing interdisciplinary merging of technologies has
led to milestones in the eld of advanced sportswear
systems. Such systems are designed to assist athletes
to reach their desired tness and performance goals by
helping to create an optimal micro- and macroenviron-
ment and/or physiological state for comfort, to facilitate
best performances by providing real-time information on
the environment and state athlete status (770). Of note,
most sensors for monitoring athletic training loads and
sports performance have been driven by technological
advances in other disciplines, mainly clinical medicine
(e.g., continuous glucose monitoring systems) and the
military (770).
The concept of wearable sensors in sport is broad
and includes any information gathering system that an
athlete can wear/carry while participating in normal train-
ing/competition environments. Wearables can be worn
by athletes on their person, in their clothing, or within
their equipment and incorporate sensors, a microproc-
essor, and a form of communication unit that enables
connectivity within a personal area network (PAN) where
a smartphone or other appliance stores data and oper-
ates as a gateway with connectivity to the internet (771).
Sensors and devices for athletes must be small and
light, as well as exible, durable, and impact resistant.
Perhaps most importantly, any device should not neg-
atively affect normal range of motion in an athletes
chosen discipline. Simultaneously, wearables must
produce precise measurements of biometrics like
motion, heart rate, blood pressure, respiration rate,
oxygen kinetics, blood, saliva, and sweat markers, and
impact forces (772).
During the past decade, there has been a global
explosion in wearables in sport and other health-related
elds (772) underpinned by rapid developments in
smart technologiessuch as articial intelligence (AI)
and machine learning. These technologies rely on sen-
sor systems that collect, process, and transmit relevant
data (such as biomarkers and other training/competition
indexes) that are crucial to evaluate an athletescondi-
tion and maximize performance (771). In the sports set-
ting, these platforms have several objectives: 1)to
gather valid, reliable, high-quality, data-rich sensory in-
formation from athletes in training/competition environ-
ments; 2) to apply sophisticated analytics methods to
identify patterns for determining athlete health and
training/competition status; 3)toobtainreal-timedata
regarding training-related metrics (i.e., training varia-
bles, sleep quality, diet) while assisting athletes/
coaches to manage a range of performance variables
and outcomes to detect early signs and symptoms of
overtraining (657); and 4)toestablishnovelperform-
ance outcomes that supplement subjective, manually
collected data and coach-based feedback with auto-
mated, objective data from devices. Data obtained
from wearables can furthermore be leveraged to learn
new techniques and provide real-time feedback on
this process (770). For example, Samsung developed
the Samsung SmartSuit to optimize body posture dur-
ing short-track speed skating. The suit includes ve
integrated sensors and enables real-time feedback to
the coach regarding the body position of the athlete:
in this scenario, real-time feedback is transmitted to
the athlete with instructions to modify body position to
reduce drag and optimize performance. The Dutch
short-tracker Suzanne Schulting used this suit to pre-
pare for the Olympic Games in 2018 and became the
rst Dutch Olympic gold medal winner for 1,000 m, a
feat she repeated in 2022.
Training load for sport performance encompasses both
external and internal dimensions, with external training
loads representing the physical work performed during a
training session and internal loads being the associated
biochemical and biomechanical stress responses. With
regard to athlete monitoring, wearables provide informa-
tion over four broad domains (771,773): 1) internal load,
representing the psychophysiological responses to a
given external load and typically determined by measure-
ment of heart rate, oxygen uptake, blood lactate, and rat-
ings of perceived exertion; 2) external load, determined
by the physical demands associated with a given stimu-
lus, monitored by global navigation satellite systems, iner-
tial measurement units, or linear/angular transducers that
provide measures related to distance covered at certain
velocities, acceleration/deceleration, and change of direc-
tion forces; 3) well-being, monitored by subjective scales
FURRER ET AL.
1752 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
related to fatigue, DOMS, stress, quality of sleep, or mode
state; and 4) readiness, assessed by measures of heart
rate variability, heart rate recovery, and variations in the
test results of selected neuromuscular or submaximal/
maximal test protocols. As long-term improvements in
training adaptation and performance capacity are ulti-
mately the result of an athletescumulativeinternalload
over multiple acute work bouts, the measurement of inter-
nal load and the factors inuencing these outcomes is of
paramount importance for the coach and athlete. A knowl-
edge of the relationships between internal and external
training loads has the potential to enhance training pre-
scription, periodization, and athlete management through
a detailed assessment of training delity and efcacy
(773). Whether such information will provide coaches with
an objective framework for evidence-based decisions
remains to be determined. Moreover, potential issues with
accuracy, dependability, or data security and privacy will
have to be considered and solutions proposed that are
acceptable for all stakeholders (774).
5.5. The Application of -OmicsTechnologies to
Exercise Biology
In the past quarter century, the eld of exercise biology
has evolved to include sophisticated analysis of multi-tis-
sues and organs through the application of established
techniques already employed in other disciplines, as
well as various -omics platforms to complement classic
approaches (775,776). Such inquiry has provided both a
greater understanding of the biological bases of the
health benets of exercise (38) as well as knowledge of
muscle bioenergetics and adaptation to training in rec-
reationally active individuals and subelite athletes (55,
160,777). The application of molecular techniques to
exercise biology has provided novel insights into the
complexity and breadth of intracellular signaling net-
works involved in response to both endurance- and re-
sistance-based exercise (55,161,777). The recent
explosion in global -omics technologies in the exercise
sciences has also provided new opportunities to map
the complexity and interconnectedness of biological
networks underlying the tissue-specic responses and
systemic benets of exercise training (161,455,586,
778782). A sportomicsapproach (the use of -omics
sciences in sports) has even been proposed to comple-
ment existing methods of studying and monitoring an
athletes state of fatigue and physical performance and
aid in talent identication (783,784). There have been
in-depth and integrated multiomics proling efforts of
the response to acute exercise in subclinical populations
(785), along with longitudinal big dataapproaches to
develop prediction models for biomarkers for precision
medicine (786). However, to date, few studies have
been undertaken in elite athletic cohorts (787).
Although there is some evidence to suggest that a com-
bined -omics solution will greatly facilitate discovery of the
genetic and non-genetic inuences on sporting perform-
ance, training response, injury predisposition, and other
potential determinants of successful human performance
(788), large-scale, collaborative efforts involving well-dened
phenotypic cohorts will be essential for major progress to
be made in the eld of elite sport performance (782).
Indeed, integration of data from multiple -omics approaches
will require large sample sizes, big data sets, and expertise
in computational biology to resolve the complex biology
associated with the diverse exercise responses to endur-
ance- and resistance-based training regimens. This will ne-
cessitate collaborative efforts between multiple research
teams using common procedures and experimental
protocols to execute multicenter exercise/lifestyle
intervention trials with the goal of collecting sufcient
functional and molecular data to further elucidate the
mechanisms responsible for adaptive response to var-
ious exercise training regimes. Issues of data privacy
and accessibility will have to be considered.
Such an approach is already underway: the Molecular
Transducers of Physical Activity Consortium (MoTrPAC)
is a multicenter study on the effects of two different
forms of exercise (endurance and resistance training)
across individuals of different ages and sexes as well as
sedentary and well-trained individuals (782). There are
two main aims of MoTrPAC: the rst is to study the
response to exercise at the whole body and cellular lev-
els and to identify the molecular underpinnings that
might be responsible for the adaptive process and varia-
tion among individuals. The second aim is to deliver
a map of the biological molecules and pathways under-
lying the systemic effects of acute and chronic exercise
(454,782). Ultimately, the knowledge gained from
MoTrPAC and other similar large-scale undertakings
(e.g., the Wu Tsai Human Performance Alliance) will give
biological science researchers and health professionals
the insights to develop personalized training protocols to
maximize performance and/or health benets based on
the unique molecular signatures and specictargets
identied. Notwithstanding the increased knowledge that
will accrue from a better understanding of these sophisti-
cated biological processes and pathways, advances in
training techniques for achieving new limits in human ath-
letic performance have rarely had their origins in science.
Part of the reason that sport science and exercise biology
has failed to inform training practices stems from a reluc-
tance of coaches to modify their methods, many of which
have been nurtured and perfected over decades and are
rmly entrenched as coaching lore.Donating tissue
samples for exercise biologists to gain mechanistic
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1753
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
insights into various training protocols has also met with
limited success. As such, knowledge of training methods
to enhance elite sport performance has traditionally
evolved by way of trial-and-error observations of a few
pioneering coaches and their athletes, with exercise sci-
entists playing post hocroles attempting to explain the
underpinning biological mechanisms (99). Although major
breakthroughs in the knowledge of how exercise acti-
vates numerous cellular, molecular, and biochemical
pathways have been witnessed, direct evidence linking
such effects to specic performance outcomes and
understanding how these effects exert their benets in
different athletic populations remains elusive and a chal-
lenge for future research. To do so, exercise biologists
who investigate training adaptation and elite athletic per-
formance will have to integrate information pertaining to
an athletes genetic and epigenetic background with tis-
sue-specic gene expression, proteome, and metabolo-
mic proles to predict potential improvements in strength,
aerobic capacity, and other traits necessary for elite
performance.
6. SUMMARY, CONCLUSIONS, AND
PERSPECTIVES
We stand on the shoulders of giants who have unraveled
seminal and fundamental aspects of muscle biology and
metabolism (789). Today, human studies are boosted by
technological and conceptual advances that enable the
investigation of molecular mechanisms, cellular functions,
multicellular dynamics, as well as inter-tissue and inter-
organ cross talk in an unprecedented manner. For exam-
ple, the identication of myokines resulted in the new de-
nition of skeletal muscle tissue as an endocrine organ, in
fact the largest in our body (790,791). Similarly, metabolism
of kynurenine (792) and excess ketone bodies in hyperke-
tonemia (575) in skeletal muscle, boosted by exercise,
imply a role for this tissue in the detoxication of dysregu-
lated endogenous metabolites, analogous to xenobiotic
detoxication in the liver. scRNA-seq has yielded novel
insights into the cellular composition of muscle, including
the identication of previously unidentied cell types and
the rst analyses of multicellular dynamics and interactions
(632634). Likewise, single-nucleus RNA-seq (snRNA-seq)
reveals a hitherto unsuspected coordination between
nucleiinthesamesyncytialmyober but also a surprising
heterogeneity and subspecication of transcriptional pro-
grams that extends beyond the classically dened
synaptic, extrasynaptic, and myotendinous nuclear
populations (793795), combined with insights into
protein transfer between nuclei (650), RNA transport
in the muscle ber (651), and even the movement of
myonuclei in speciccontexts(649). Exon skipping,
CRISPR-Cas9-based approaches, and adeno-associ-
ated viral (AAV) vectors have been applied in different
settings and pathologies, including muscle diseases,
in the preclinical setting, such as therapy of Duchenne
muscular dystrophy (796). AAV-based gene therapy is
now also used clinically, for example, in spinal muscu-
lar atrophy patients (797).Finally,progressisbeing
made in the recognition of our increasingly sedentary
lifestyleasamajorriskfactorforchronicmetabolic
diseases, the prescription of exercise-based interven-
tions in the general population for preventing and/or
treating an ever-increasing number of widespread
conditions, and establishing physical activity as a cor-
nerstone in medical practice and public health (26,
798,799).
In elite sport, world records continue to be broken
across a wide range of events. Performance improve-
ments or declines depend on many factors, including
technology, sports science, support for a particular
sport, talent identication, investment of time, and effec-
tiveness of training protocols (800802). Advances in
elite performance can also vary between individual
events in one sport, as is the case for Olympic swimming
competitions, with strong trends for improvement in
some strokes and a relative plateau in others (803). With
an increasing number of former Olympians and elite ath-
letes now participating in masters competitions, age-
group records are being surpassed and seemingly unat-
tainable performances recorded such as the sub-3 hour
marathon by a 70-yr-old man (804). Although big data
approaches will be facilitated by the rapidly evolving
tracking systems and wearables technologies combined
with machine learning, AI, and other analysis methods
(66,805,806), it is questionable as to what extent train-
ing practices of elite athletes have been facilitated by
any major laboratory-based scientic breakthroughs to
date. Indeed, despite our greater understanding of
some of the mechanistic underpinnings of muscle plas-
ticity and exercise adaptation, the upper limits of adapta-
tion remain poorly studied (18). Even though numerous
potential regulatory and functional key players have
been identied, we do not know whether the picture is
complete, how these factors are activated and engaged,
how different pathways are integrated, or how the regu-
latory and functional outcome is specied, orchestrated,
and coordinated. Moreover, the complexity of the appa-
rent regulatory and functional redundancies, contingen-
cies, alternatives, and adaptive mechanisms that ensure
robust regulation of muscle plasticity as one of the most
fundamental aspects of human life and evolution
remains enigmatic (18). Our insights into the regulation
of muscle plasticity in response to endurance-based
training stimuli far surpasses that of resistance-based
exercise, in part because of the availability of more
FURRER ET AL.
1754 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
robust or at least more commonly used animal models
and protocols that are more physiological and translat-
able for the former training practice (807,808).
Compared to the muscle and whole body responses to
an acute bout of exercise, the mechanistic aspects of
chronic training-induced plasticity are less well investi-
gated and understood. Similarly, the mechanistic under-
standing of the molecular bases of reduced training
and/or detraining, as well as of retraining and muscle
memory, remains rudimentary (433). Finally, we have lim-
ited knowledge of the dynamic multidirectional cross
talk between muscle and other cell types, within and
beyond muscle tissue, which is instrumental for
adequate muscle function and delays a better delinea-
tion of training adaptation. For example, studies of the
motor unit, the unity of muscle ber and motor neuron,
should include sensory-motor circuits in the spinal cord
and supraspinal systems, brain stem neurons with de-
scending axons, and brain regions that are involved in
locomotor control and integrate sensory feedback, all
important for muscle control, resistance training adapta-
tion, muscle memory, and other processes (329,456,
464,809).
How will we overcome these shortcomings, and what
could the future of muscle research look like? In addition
to new advances in technological possibilities, analysis
methods, and computational modeling approaches, one
aim would be stronger interaction, collaboration, and
networking between basic science, sports sciences, and
coaches as well as athletes (FIGURE 21). All these elds,
when optimally synergized, may help to obtain a better
understanding of muscle plasticity from the inactive to
the extreme (810,811). Currently, training paradigms pio-
neered by athletes tend to inform and guide research,
as was the case for interval training practiced by Paavo
Nurmi in the 1920s and Emil Zátopek in the 1950s for
middle- and long-distance running disciplines (812).
Such innovative methods are currently enjoying intense
scientic scrutiny in the rened form of HIIT, for both ath-
letic endeavors and the tness and well-being of the
general population (100). Athlete feedback is central to
understanding individualized training response, fatigue
recovery, or concurrent training design and helps to study
the mechanistic aspects that underlie these processes
for iterative optimization of training and competition
(71). Importantly, sports psychology and neurobiology
should integrate morphological, cellular, and mecha-
nistic aspects to identify the relevant circuitries,
regions, and signals involved in the control of these
factors, including the cross talk between muscle and
brain (494,813,814).
Bypassing this chain of research by jumping directly
from basic science to athletes often leads to a mismatch
between preclinical data and real-worldperformance.
For example, based on mechanistic insights and mouse
experiments with so-called exercise mimetics,some
coaches and athletes experimented with performance-
enhancing compounds, most recently AMPK and PPARb/
dactivators, without waiting for robust scientic validation
from human trials (45,46). Not only is there no evidence
for performance enhancement in humans (815), but in
somecases(e.g.,metformin,resveratrol,andrapamycin),
there may be a reduction of training adaptation (45,46).
Moreover, such compounds may have a signicant risk of
severeadverseeffectsthatmightnotberelevantinthe
time frame of application and life expectancy in rodents
but could have detrimental long-term effects in humans
(45,46). For example, prolonged, sustained activation of
AMPK could lead to a catabolic state, lactic acidosis, car-
diac hypertrophy, brain inammation, and reduced cogni-
tive abilities (45,46,816). PPARb/dligands increase the
risk of tumors in rodents when given at high doses over a
prolonged period of time, and even AMPKsactioncan
switch from tumor suppressor to tumor promoter once
cancer develops (45,46,817). It is encouraging to see
that integrative approaches are increasingly pursued at
different centers and organizations in which elite athlete
training and health management is under the same roof
as integrative research. Thus, collectively, mechanistic
understanding, implementation in training design, techno-
logical innovations, and other advances, cross-fertilized
with data from psychology, nutrition, and sleep research,
will help to further optimize athletic performance in a
safe, personalized, and evidence-based approach (818,
819). Thereby, pseudoscience, baseless claims, quick
xes,and other potentially detrimental interventions can
be minimized and information separated from misinfor-
mation (820). Finally, integrative research in exercise
should be combined with exercise medicine to further our
understanding of the immense potential of exercise-
based interventions to prevent and treat many chronic
diseases in the general population (821), along with
the use of training for injury rehabilitation in physical ther-
apy. A better understanding in the athletic setting thus
could and should inform interventions in the general popu-
lation and in patients (822). For example, studies of resil-
ience and motivation might also help to design approaches
to facilitate and improve adherence and compliance to exer-
cise training (823825). Such an approach could leverage
novel avenues such as virtual reality exergames (826,827).
Insights into exercise physiology are leveraged in disease
diagnostic, prevention, treatment, and rehabilitation, e.g.,
the use of blood pressure measurements during exercise
(exercise hypertension) to reveal undiagnosed or masked
hypertension and predict cardiovascular disease risks
(828,829). Many more areas exist in which concepts
derived from elite sports will benet non-athletes and
patients (822). In this context, exercise sociology
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1755
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
could contribute to overcome existing individual
and societal challenges (830). Hopefully, exercise
will rmly be recognized and accepted in all aspects
of clinical practice and the political and societal
framework established to facilitate and promote an
active lifestyle.
Finally, an issue that will have to be addressed in the
eld of biological/sports science is the formulation of
hypotheses and interpretation of data and results. Too
often, papers claim to have solved an open question in
an absolute manner, even though in many cases seem-
ingly conicting, contradictory, or non-overlapping alter-
native studies exist. Often, such studies depend on
relatively small subject numbers and might have to be
interpreted in light of age, sex, training protocol, sam-
pling time, and other parameters (diet, sleep, comorbid-
ities, tness level, chronotype, and time of day of the
study) (831). Human and animal studies are guilty of this
alike. Studies in model organisms have a limited predic-
tive power for human exercise physiology but enable
Large cohort data, wearables
Individual athletes
Model organisms
Molecular biology
Prevention and treatment
Mechanisms,
Causality
Human relevance
(descriptive, correlative)
Athlete/coach
feedback
Sports psychology
Nutritional sciences
Chronobiology and sleep research
Basic muscle
science
Human exercise
physiology
Exercise is
medicine
Cell culture models,
organoids
Computational biology,
modeling
FIGURE 21. The future of exercise science for safe, evidence-based, and personalized approaches. To overcome existing hurdles and efciently
leverage the power of novel techniques and approaches, a close interaction between basic muscle research, applied human exercise physiol-
ogy, as well as athletes and coaches should be aimed for. Model organisms, cell culture, and molecular and computational biology might provide
insights into cause-effect relationships, epistasis, etiologies, and mechanisms complementing the descriptive and correlative studies in human
volunteers. Inversely, data from large cohorts that will become available because of widespread use of wearables and trackers as well as those
obtained in and based on feedback from athletes and coaches will reveal processes and pathways to be explored in mechanistic detail. The mo-
lecular athlete should furthermore be informed by sports psychology, e.g., in regard to motivation, perseverance, compliance and adherence,
nutritional sciences, chronobiology, sleep research, and other elds of relevance to training. Finally, a mutual exchange between the observa-
tions in training and those in various pathologies associated with an inactive lifestyle or inadequate muscle functionality will help to push the
boundaries of physical activity interventions in the prevention and treatment of numerous diseases. Image created with BioRender.com, with
permission.
FURRER ET AL.
1756 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
causative and mechanistic insights, which are difcult to
conduct in human trials that rely mainly on descriptive or
correlative data. Moreover, while model organisms pro-
vide access to all different types of muscle beds and
allow the analysis of whole muscles, human studies are
in most cases limited to small, single biopsies of one
muscle (typically the vastus lateralis), taken at different
times before, during, or after exercise, with considerable
variability in outcomes and hence interpretation of
results (388,446,832). It might be advisable to be aware
of possible limitations and keep an open mind vis-à-vis
alternative interpretations, study-specic bias, and sys-
tem complexity surpassing simplistic explanations. For
example, there most likely is more than one cause for
muscle fatigue (833); satellite cell recruitment might be
important for hypertrophy in some but not all cases
(658); the relative contribution of training intensity and
volume on mitochondrial function in muscle might be
very context-dependent (104,105) as is the relative con-
tribution of muscle hypertrophy to gains in strength
(834,835); and polarized or pyramidal intensity distribu-
tions might be optimal for performance enhancement
(116,132134). Even established principles,”“dogmas,
and lawsshould constantly be questioned, validated,
and rened. Unfortunately, more often than not, litera-
ture searches for such ndings disappear into a trail of
never-ending, consecutive citations. In addition, data
and the arising hypotheses must be examined under
consideration of the technical and conceptual possibil-
ities of the respective historical time period. For exam-
ple, Hennemans size principle of motor unit recruitment
(836) certainly holds up under the laboratory conditions
using the exact preparations and the methods that were
used in the 1950s to 1970s. However, the rigid view of
this principle has since been rened and modied with
novel approaches (i.e., using more complex musculoten-
don-skeletal systems, physiologically relevant range of
forces, physiological and neural stimulation as opposed
to electromyography and -physiology, or consideration
of neural drive, cortical and afferent input to each indi-
vidual motor neuron) (319,321,330,625). Moreover, it is
not clear how the size principle can accommodate mus-
cle ber type shifts in exercise (624), and recent scRNA-
seq and snRNA-seq approaches revealed a greater di-
versity in motor neuron populations than the classically
dened types based on transcriptional proles (326,
837). Thus, even though probably correct at its core, the
size principle might be oversimplied, and motor unit
recruitment and plasticity certainly warrant further study.
Similarly, even though Nobel prize worthy, August
Kroghs ideas on oxygen delivery and muscle microvas-
culature have, for some aspects, not withstood the test
of time or have been rened and altered by more mod-
ern and comprehensive methods (595). The myonuclear
domain hypothesis was postulated to account for the
syncytial nature of myobers, implying that a nucleus is
needed for adequate support for transcription and trans-
lation within a speciccellvolume,henceaxed myonu-
clear domain, in these extraordinarily large cells (838).
Accordingly, satellite cell recruitment would be needed
to provide additional nuclei in ber hypertrophy (839).
However, the myonuclear domain hypothesis fails to
provide an adequate and complete explanation for sev-
eral observations. First, removal of myonuclei in atro-
phy is controversial and not observed consistently
(839843). Second, the myonuclear domain has high
exibility and scales with body size, ber type, mito-
chondrial activity, ber hypertrophy, and other param-
eters (838,844846). Third, even though the spacing
between most myonuclei is roughly even, regional
differences exist, most notably in the tight clusters of
three to ve subsynaptic myonuclei at the NMJ and
similar clusters at the myotendinous junction, for both
of which one nucleus should theoretically be suf-
cient to serve the respective cytoplasmic domain
(468). Ample evidence exists of intracellular move-
ment of myonuclei (847), as in postexercise ber
repair (649), exchange of proteins between nuclei
(650), and microtubule-mediated transport of ribonu-
cleoproteins and RNAs within the myober (651), all of
which imply a highly plastic system transcending a
more rigid denition of myonuclear domains. Finally,
in cells that rival myobers in terms of size or length,
such as certain motor neurons with an axonal length
of >1m(848) compared with some of the longest
muscle bers in the human musculus sartorius reach-
ing the length of 60 cm (849), one nucleus, located
asymmetrically in the cell body in the spinal cord,
seems sufcient to provide transcripts for the whole
cell. Thus, the evolutionary pressure for and physio-
logical function of the syncytial nature of myobers
remain largely mysterious.
These are just a few examples to illustrate that in exer-
cise biology and sport science, as in any place of scien-
tic and social discourse (850852), we should have
passionate arguments but remain fair, civil, agnostic,
and open minded, carefully consider alternative results
and hypotheses, and constantly challenge, validate, and
rene (or refute!) seemingly establishedprinciples. In
the nal analyses, modern sports science and exercise
biology offer numerous opportunities to assist elite ath-
letes to rene training methods, optimize adaptation,
stay healthy and injury free, achieve their desired phy-
sique, and ght against fatigue factors that limit success-
ful performance. The accomplishments of elite athletes
will continue to entertain and amaze us, as science
attempts to catch up and explain the biological bases of
such feats.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1757
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
GLOSSARY
AMPK AMP-dependent protein kinase
ATP Adenosine triphosphate
CSA Cross-sectional area
DOMS Delayed-onset muscle soreness
ECM Extracellular matrix
EPO Erythropoietin
Hb Hemoglobin
HIF-1aHypoxia-inducible factor 1a
HIIT High-intensity interval training
IL Interleukin
MPS Muscle protein synthesis
mTOR Mammalian target of rapamycin
mTORC1 mTOR complex 1
NMJ Neuromuscular junction
OXPHOS Oxidative phosphorylation
PGC-1aPeroxisome proliferator-activated receptor ccoac-
tivator 1a(gene name PPARGC1A)
1RM One-repetition maximum
RYR Ryanodine receptor
SERCA Sarcoplasmic/endoplasmic reticulum Ca2
1
-
ATPase
SNP Single-nucleotide polymorphism
SR Sarcoplasmic reticulum
VEGF Vascular epithelial growth factor
_
VO2max Maximal oxygen uptake
CORRESPONDENCE
C. Handschin (christoph.handschin@unibas.ch); R. Furrer
(regula.furrer@unibas.ch); J. A. Hawley ( john.hawley@acu.
edu.au).
ACKNOWLEDGMENTS
The authors thank the members of their research groups for
helpful discussions and comments regarding this manuscript.
We also thank the coaches and athletes for sharing training
plans. Because of space limitations, we have been unable to
include all of the pertinent and original work by some of our
peers, for which we apologize in advance. The gures were
created with BioRender.com. Images were obtained from pub-
lic repositories and acknowledged in the respective gure
legends.
GRANTS
Work in the laboratory of R.F. and C.H. related to this article
was supported by the Swiss National Science Foundation
(310030_184832), the European Research Council (616830-
MUSCLE_NET), Innosuisse (44112.1 IP-LS), the Swiss Society
for Research on Muscle Diseases, the Jain Foundation, the
Biozentrum, and the University of Basel.
DISCLOSURES
C.H. is an associate editor of Physiological Reviews and was
not involved in and did not have access to information regard-
ing the peer-review process or nal disposition of this article.
An alternate editor oversaw the peer-review and decision-
making process for this article. None of the other authors has
any conicts of interest, nancial or otherwise, to disclose.
AUTHOR CONTRIBUTIONS
R.F., J.A.H., and C.H. prepared gures; drafted manuscript;
edited and revised manuscript; and approved the nal version
of the manuscript.
REFERENCES
1. Marino FE, Sibson BE, Lieberman DE.The evolution of human fatigue
resistance. J Comp Physiol B 192: 411422, 2022. doi:10.1007/
s00360-022-01439-4.
2. Bramble DM, Lieberman DE.Endurance running and the evolution
of Homo.Nature 432: 345352, 2004. doi:10.1038/nature03052.
3. Hunter P.The evolution of human endurance: research on the biol-
ogy of extreme endurance gives insights into its evolution in humans
and animals. EMBO Rep 20: e49396, 2019. doi:10.15252/embr.
201949396.
4. Kim DS, Wheeler MT, Ashley EA.The genetics of human perform-
ance. Nat Rev Genet 23: 4054, 2022. doi:10.1038/s41576-021-
00400-5.
5. Liebenberg L.The relevance of persistence hunting to human evo-
lution. J Hum Evol 55: 11561159, 2008. doi:10.1016/j.jhevol.2008.
07.004.
6. Pontzer H.Economy and endurance in human evolution. Curr Biol
27: R613R621, 2017. doi:10.1016/j.cub.2017.05.031.
7. Kiely J, Collins DJ.Uniqueness of human running coordination: the
integration of modern and ancient evolutionary innovations. Front
Psychol 7: 262, 2016. doi:10.3389/fpsyg.2016.00262.
8. Lieberman DE.Human locomotion and heat loss: an evolutionary
perspective. Compr Physiol 5: 99117, 2015. doi:10.1002/cphy.
c140011.
9. HawleyJA,HargreavesM,JoynerMJ,ZierathJR.Integrative biology
of exercise. Cell 159: 738749, 2014. doi:10.1016/j.cell.2014.10.029.
10. Mattson MP.Evolutionary aspects of human exerciseborn to run
purposefully. Ageing Res Rev 11: 347352, 2012. doi:10.1016/j.
arr.2012.01.007.
11. Neel JV.Diabetes mellitus: a thriftygenotype rendered detrimen-
tal by progress?Am J Hum Genet 14: 353362, 1962.
12. Booth FW, Chakravarthy MV, Gordon SE, Spangenburg EE.Waging
war on physical inactivity: using modern molecular ammunition
against an ancient enemy. J Appl Physiol (1985) 93: 330, 2002.
doi:10.1152/japplphysiol.00073.2002.
13. Rodriguez J, Vernus B, Chelh I, Cassar-Malek I, Gabillard JC, Hadj
Sassi A, Seiliez I, Picard B, Bonnieu A.Myostatin and the skeletal
FURRER ET AL.
1758 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
muscle atrophy and hypertrophy signaling pathways. Cell Mol Life
Sci 71: 43614371, 2014. doi:10.1007/s00018-014-1689-x.
14. Bellinge RH, Liberles DA, Iaschi SP, OBrien PA, Tay GK.Myostatin
and its implications on animal breeding: a review. Anim Genet 36:
16, 2005. doi:10.1111/j.1365-2052.2004.01229.x.
15. Lee SJ.Targeting the myostatin signaling pathway to treat muscle
loss and metabolic dysfunction. JClinInvest131: e148372, 2021.
doi:10.1172/JCI148372.
16. Rodgers BD, Garikipati DK.Clinical, agricultural, and evolutionary
biology of myostatin: a comparative review. Endocr Rev 29: 513
534, 2008. doi:10.1210/er.2008-0003.
17. Eliasson P, Andersson T, Kulas J, Seemann P, Aspenberg P.
Myostatin in tendon maintenance and repair. Growth Factors 27:
247254, 2009. doi:10.1080/08977190903052539.
18. Lavin KM, Coen PM, Baptista LC, Bell MB, Drummer D, Harper SA,
Lixandrão ME, McAdam JS, OBryan SM, Ramos S, Roberts LM,
Vega RB, Goodpaster BH, Bamman MM, Buford TW.State of knowl-
edge on molecular adaptations to exercise in humans: historical
perspectives and future directions. Compr Physiol 12: 31933279,
2022. doi:10.1002/cphy.c200033.
19. King AJ, Burke LM, Halson SL, Hawley JA.The challenge of main-
taining metabolic health during a global pandemic. Sports Med 50:
12331241, 2020. doi:10.1007/s40279-020-01295-8.
20. Jakobsson J, Cotgreave I, Furberg M, Arnberg N, Svensson M.
Potential physiological and cellular mechanisms of exercise that
decrease the risk of severe complications and mortality following
SARS-CoV-2 infection. Sports (Basel) 9: 121, 2021. doi:10.3390/
sports9090121.
21. Bull FC, Al-Ansari SS, Biddle S, Borodulin K, Buman MP, Cardon G,
Carty C, Chaput JP, Chastin S, Chou R, Dempsey PC, DiPietro L,
Ekelund U, Firth J, Friedenreich CM, Garcia L, Gichu M, Jago R,
Katzmarzyk PT, Lambert E, Leitzmann M, Milton K, Ortega FB,
Ranasinghe C, Stamatakis E, Tiedemann A, Troiano RP, van der
Ploeg HP, Wari V, Willumsen JF.World Health Organization 2020
guidelines on physical activity and sedentary behaviour. Br J Sports
Med 54: 14511462, 2020. doi:10.1136/bjsports-2020-102955.
22. McKay AK, Stellingwerff T, Smith ES, Martin DT, Mujika I, Goosey-
Tolfrey VL, Sheppard J, Burke LM.Dening training and performance
caliber: a participant classication framework. Int J Sports Physiol
Perform 17: 317331, 2022. doi:10.1123/ijspp.2021-0451.
23. Levine BD.VO2max: what do we know, and what do we still need
to know? J Physiol 586: 2534, 2008. doi:10.1113/jphysiol.2007.
147629.
24. Costill DL.Metabolic responses during distance running. JAppl
Physiol 28: 251255, 1970. doi:10.1152/jappl.1970.28.3.251.
25. Costill DL, Fink WJ, Pollock ML.Muscle ber composition and
enzyme activities of elite distance runners. Med Sci Sports 8: 96
100, 1976.
26. MyersJ,PrakashM,FroelicherV,DoD,PartingtonS,Atwood
JE.Exercise capacity and mortality among men referred for
exercise testing. NEnglJMed346: 793801, 2002. doi:10.1056/
NEJMoa011858.
27. Strasser B, Burtscher M.Survival of the ttest: VO2max, a key pre-
dictor of longevity? Front Biosci (Landmark Ed) 23: 15051516,
2018. doi:10.2741/4657.
28. Kodama S, Saito K, Tanaka S, Maki M, Yachi Y, Asumi M, Sugawara
A, Totsuka K, Shimano H, Ohashi Y, Yamada N, Sone H.
Cardiorespiratory tness as a quantitative predictor of all-cause mor-
tality and cardiovascular events in healthy men and women: a meta-
analysis. JAMA 301: 20242035, 2009. doi:10.1001/jama.2009.681.
29. Lavie CJ, Ozemek C, Carbone S, Katzmarzyk PT, Blair SN.
Sedentary behavior, exercise, and cardiovascular health. Circ Res
124: 799815, 2019. doi:10.1161/CIRCRESAHA.118.312669.
30. Katzmarzyk PT, Ross R, Blair SN, Despr
es JP.Should we target
increased physical activity or less sedentary behavior in the battle
against cardiovascular disease risk development? Atherosclerosis
311: 107115, 2020. doi:10.1016/j.atherosclerosis.2020.07.010.
31. Srikanthan P, Karlamangla AS.Muscle mass index as a predictor of
longevity in older adults. Am J Med 127: 547553, 2014.
doi:10.1016/j.amjmed.2014.02.007.
32. Linge J, Petersson M, Forsgren MF, Sanyal AJ, Dahlqvist Leinhard
O.Adverse muscle composition predicts all-cause mortality in the
UK Biobank imaging study. J Cachexia Sarcopenia M uscle 12:
15131526, 2021. doi:10.1002/jcsm.12834.
33. Liu M, Zhang Z, Zhou C, Ye Z, He P, Zhang Y, Li H, Liu C, Qin X.
Predicted fat mass and lean mass in relation to all-cause and
cause-specic mortality. J Cachexia Sarcopenia Muscle 13: 1064
1075, 2022. doi:10.1002/jcsm.12921.
34. Rantanen T, Masaki K, He Q, Ross GW, Willcox BJ, White L.Midlife
muscle strength and human longevity up to age 100 years: a 44-
year prospective study among a decedent cohort. Age (Dordr) 34:
563570, 2012. doi:10.1007/s11357-011-9256-y.
35. Sayer AA, Kirkwood TB.Grip strength and mortality: a biomarker of
ageing? Lancet 386: 226227, 2015. doi:10.1016/S0140-6736(14)
62349-7.
36. Li R, Xia J, Zhang XI, Gathirua-Mwangi WG, Guo J, Li Y, McKenzie S,
Song Y.Associations of muscle mass and strength with all-cause
mortality among US older adults. Med Sci Sports Exerc 50: 458
467, 2018. doi:10.1249/MSS.0000000000001448.
37. Guan Y, Yan Z.Molecular mechanisms of exercise and healthspan.
Cells 11: 872, 2022. doi:10.3390/cells11050872.
38. Hargreaves M.Exercise and health: historical perspectives and
new insights. J Appl Physiol (1985) 131: 575588, 2021. doi:10.1152/
japplphysiol.00242.2021.
39. McGee SL, Hargreaves M.Exercise adaptations: molecular
mechanisms and potential targets for therapeutic benet. Nat
Rev Endocrinol 16: 495505, 2020. doi:10.1038/s41574-020-
0377-1.
40. Lim C, Nunes EA, Currier BS, McLeod JC, Thomas AC, Phillips SM.An
evidence-based narrative review of mechanisms of resistance exer-
cise-induced human skeletal muscle hypertrophy. Med Sci Sports
Exerc 54: 15461559, 2022. doi:10.1249/MSS.0000000000002929.
41. Egan B, Sharples AP.Molecular responses to acute exercise and
their relevance for adaptations in skeletal muscle to exercise train-
ing. Physiol Rev. In Press. doi:10.1152/physrev.00054.2021.
42. Booth FW, Roberts CK, Thyfault JP, Ruegsegger GN, Toedebusch
RG.Role of inactivity in chronic diseases: evolutionary insight and
pathophysiological mechanisms. Physiol Rev 97: 13511402, 2017.
doi:10.1152/physrev.00019.2016.
43. Handschin C, Spiegelman BM.The role of exercise and PGC1alpha
in inammation and chronic disease. Nature 454: 463469, 2008.
doi:10.1038/nature07206.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1759
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
44. Tremblay MS, Colley RC, Saunders TJ, Healy GN, Owen N.
Physiological and health implications of a sedentary lifestyle. Appl
Physiol Nutr Metab 35: 725740, 2010. doi:10.1139/H10-079.
45. Handschin C.Caloric restriction and exercise mimetics: ready for
prime time? Pharmacol Res 103: 158166, 2016. doi:10.1016/j.
phrs.2015.11.009.
46. Hawley JA, Joyner MJ, Green DJ.Mimicking exercise: what matters
most and where to next? J Physiol 599: 791802, 2021. doi:10.1113/
JP278761.
47. Weihrauch M, Handschin C.Pharmacological targeting of exercise
adaptations in skeletal muscle: benets and pitfalls. Biochem
Pharmacol 147: 211220, 2018. doi:10.1016/j.bcp.2017.10.006.
48. Berthelot G, Sedeaud A, Marck A, Antero-Jacquemin J, Schipman J,
Sauli
ere G, Marc A, Desgorces FD, Toussaint JF.Has athletic per-
formance reached its peak? Sports Med 45: 12631271, 2015.
doi:10.1007/s40279-015-0347-2.
49. Marck A, Antero J, Berthelot G, Johnson S, Sedeaud A, Leroy A,
Marc A, Spedding M, Di Meglio JM, Toussaint JF.Age-related upper
limits in physical performances. J Gerontol A Biol Sci Med Sci 74:
591599, 2019. doi:10.1093/gerona/gly165.
50. Godsey B.Comparing and forecasting performances in different
events of athletics using a probabilistic model. J Quant Anal Sports
8: 2, 2012. doi:10.1515/1559-0410.1434.
51. Hoogkamer W, Snyder KL, Arellano CJ.Modeling the benets of co-
operative drafting: is there an optimal strategy to facilitate a sub-2-
hour marathon performance? Sports Med 48: 28592867, 2018.
doi:10.1007/s40279-018-0991-4.
52. Hoogkamer W, Snyder KL, Arellano CJ.Reecting on Eliud
Kipchoges marathon world record: an update to our model of co-
operative drafting and its potential for a sub-2-hour performance.
Sports Med 49: 167170, 2019. doi:10.1007/s40279-019-01056-2.
53. Jones AM, Kirby BS, Clark IE, Rice HM, Fulkerson E, Wylie LJ,
Wilkerson DP, Vanhatalo A, Wilkins BW.Physiological demands of
running at 2-hour marathon race pace. J Appl Physiol (1985) 130:
369379, 2021. doi:10.1152/japplphysiol.00647.2020.
54. Joyner MJ, Hunter SK, Lucia A, Jones AM.Physiology and fast mar-
athons. J Appl Physiol (1985) 128: 10651068, 2020. doi:10.1152/
japplphysiol.00793.2019.
55. Hawley JA, Lundby C, Cotter JD, Burke LM.Maximizing cellular ad-
aptation to endurance exercise in skeletal muscle. Cell Metab 27:
962976, 2018. doi:10.1016/j.cmet.2018.04.014.
56. Snyder KL, Hoogkamer W, Triska C, Taboga P, Arellano CJ, Kram R.
Effects of course design (curves and elevation undulations) on mar-
athon running performance: a comparison of Breaking 2 in Monza
and the INEOS 1:59 Challenge in Vienna. J Sports Sci 39: 754759,
2021. doi:10.1080/02640414.2020.1843820.
57. Akkari A, Machin D, Tanaka H.Greater progression of athletic per-
formance in older Masters athletes. Age Ageing 44: 683686,
2015.
58. Lippi G, BanG, Favaloro EJ, Rittweger J, Maffulli N.Updates on
improvement of human athletic performance: focus on world
records in athletics. Br Med Bull 87: 715, 2008. doi:10.1093/bmb/
ldn029.
59. Lepers R, Stapley PJ.Master athletes are extending the limits of
human endurance. Front Physiol 7: 613, 2016. doi:10.3389/
fphys.2016.00613.
60. Haake SJ.The impact of technology on sporting performance in
Olympic sports. JSportsSci27: 14211431, 2009. doi:10.1080/
02640410903062019.
61. OConnor LM, Vozenilek JA.Is it the athlete or the equipment?
An analysis of the top swim performances from 1990 to 2010.
JStrengthCondRes25: 32393241, 2011. doi:10.1519/JSC.
0b013e3182392c5f.
62. Malizia F, Blocken B.Bicycle aerodynamics: history, state-of-the-art
and future perspectives. J Wind Eng Indust Aerodyn 200: 104134,
2020. doi:10.1016/j.weia.2020.104134.
63. Kuper GH, Sterken E.Endurance in speed skating: the develop-
ment of world records. Eur J Oper Res 148: 293301, 2003.
doi:10.1016/S0377-2217(02)00685-9.
64. Dyer B.A pragmatic approach to resolving technological unfair-
ness: the case of NikesVapory and Alphay running footwear.
Sports Med Open 6: 21, 2020. doi:10.1186/s40798-020-00250-1.
65. Hunter I, McLeod A, Valentine D, Low T, Ward J, Hager R.Running
economy, mechanics, and marathon racing shoes. JSportsSci37:
23672373, 2019. doi:10.1080/02640414.2019.1633837.
66. Emig T, Peltonen J.Human running performance from real-world
big data. Nat Commun 11: 4936, 2020. doi:10.1038/s41467-020-
18737-6.
67. Joyner MJ.Genetic approaches for sports performance: how far
away are we? Sports Med 49: 199204, 2019. doi:10.1007/s40279-
019-01164-z.
68. Williams AG, Wackerhage H, Day SH.Genetic testing for sports per-
formance, responses to training and injury risk: practical and ethical
considerations. Med Sport Sci 61: 105119, 2016. doi:10.1159/
000445244.
69. Hecksteden A, Kraushaar J, Scharhag-Rosenberger F, Theisen D,
Senn S, Meyer T.Individual response to exercise traininga statistical
perspective. J Appl Physiol (1985) 118: 14501459, 2015. doi:10.1152/
japplphysiol.00714.2014.
70. Larsen HB, Sheel AW.The Kenyan runners. Scand J Med Sci
Sports 25: 110118, 2015. doi:10.1111/sms.12573.
71. Etxebarria N, Mujika I, Pyne DB.Training and competition readiness
in triathlon. Sports (Basel) 7: 101, 2019. doi:10.3390/sports7050101.
72. Guest NS, Horne J, Vanderhout SM, El-Sohemy A.Sport nutrige-
nomics: personalized nutrition for athletic performance. Front Nutr
6: 8, 2019. doi:10.3389/fnut.2019.00008.
73. Kerksick CM, Wilborn CD, Roberts MD, Smith-Ryan A, Kleiner SM,
J
ager R, Collins R, Cooke M, Davis JN, Galvan E, Greenwood M,
Lowery LM, Wildman R, Antonio J, Kreider RB.ISSN exercise &
sports nutrition review update: research & recommendations. J
Int Soc Sports Nutr 15: 38, 2018. doi:10.1186/s12970-018-0242-y.
74. Coyle EF, Burton HM, Satiroglu R.Inactivity causes resistance to
improvements in metabolism after exercise. Exerc Sport Sci Rev
50: 8188, 2022. doi:10.1249/JES.0000000000000280.
75. Foster C, Barroso R, Beneke R, Bok D, Boullosa D, Casado A, Chamari
K, Cortis C, Koning J, Fusco A, Haugen T, Lucía A, Mujika I, Pyne D,
Rodríguez-Marroyo JA, Sandbakk O, Seiler S.How to succeed as an
athlete: what we know, what we need to know. Int J Sports Physiol
Perform 17: 333334, 2022. doi:10.1123/ijspp.2021-0541.
76. Lepers R, Stapley PJ, Cattagni T, Gremeaux V, Knechtle B.Limits in
endurance performance of octogenarian athletes. J Appl Physiol
(1985) 114: 829, 2013. doi:10.1152/japplphysiol.00038.2013.
FURRER ET AL.
1760 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
77. Lepers R, Stapley PJ, Cattagni T.Centenarian athletes: examples of
ultimate human performance? Age Ageing 45: 732736, 2016.
doi:10.1093/ageing/afw111.
78. Booth FW, Laye MJ.Lack of adequate appreciation of physical
exercises complexities can pre-empt appropriate design and inter-
pretation in scienticdiscovery.J Physiol 587: 55275539, 2009.
doi:10.1113/jphysiol.2009.179507.
79. Hawley JA.Adaptations of skeletal muscle to prolonged, intense
endurance training. Clin Exp Pharmacol Physiol 29: 218222,
2002. doi:10.1046/j.1440-1681.2002.03623.x.
80. Joyner MJ, Coyle EF.Endurance exercise performance: the physiol-
ogy of champions. J Physiol 586: 3544, 2008. doi:10.1113/jphysiol.
2007.143834.
81. Burke LM, Hawley JA.Swifter, higher, stronger: whatsonthe
menu? Science 362: 781787, 2018. doi:10.1126/science.aau2093.
82. Jones CM, Grifths PC, Mellalieu SD.Training load and fatigue
marker associations with injury and illness: a systematic review of
longitudinal studies. Sports Med 47: 943974, 2017. doi:10.1007/
s40279-016-0619-5.
83. Sarzynski MA, Bouchard C.World-class athletic performance and
genetic endowment. Nat Metab 2: 796798, 2020. doi:10.1038/
s42255-020-0233-6.
84. Dent JR, Stocks B, Campelj DG, Philp A.Transient changes to met-
abolic homeostasis initiate mitochondrial adaptation to endur-
ance exercise. Semin Cell Dev Biol 143: 316, 2023. doi:10.1016/
j.semcdb.2022.03.022.
85. Camera DM, Smiles WJ, Hawley JA.Exercise-induced skeletal muscle
signaling pathways and human athletic performance. Free Radic Biol
Med 98: 131143, 2016. doi:10.1016/j.freeradbiomed.2016.02.007.
86. Coffey VG, Hawley JA.The molecular bases of training adapta-
tion. Sports Med 37: 737763, 2007. doi:10.2165/00007256-
200737090-00001.
87. Egan B, Zierath JR.Exercise metabolism and the molecular regula-
tion of skeletal muscle adaptation. Cell Metab 17: 162184, 2013.
doi:10.1016/j.cmet.2012.12.012.
88. Coffey VG, Hawley JA.Concurrent exercise training: do opposites
distract? J Physiol 595: 28832896, 2017. doi:10.1113/JP272270.
89. Callahan MJ, Parr EB, Snijders T, Conceição MS, Radford BE,
Timmins RG, Devlin BL, Hawley JA, Camera DM.Skeletal muscle
adaptive responses to different types of short-term exercise train-
ing and detraining in middle-age men. Med Sci Sports Exerc 53:
20232036, 2021. doi:10.1249/MSS.0000000000002684.
90. Coyle EF, Coggan AR, Hemmert MK,Ivy JL.Muscle glycogen utiliza-
tion during prolonged strenuous exercise when fed carbohydrate. J
Appl Physiol (1985) 61: 165172, 1986. doi:10.1152/jappl.1986.
61.1.165.
91. Coyle EF, Martin WH 3rd, Bloomeld SA, Lowry OH, Holloszy JO.
Effects of detraining on responses to submaximal exercise. JAppl
Physiol (1985) 59: 853859, 1985. doi:10.1152/jappl.1985.59.3.853.
92. CoyleEF,MartinWH3rd,SinacoreDR,JoynerMJ,HagbergJM,
Holloszy JO.Time course of loss of adaptations after stopping pro-
longed intense endurance training. J Appl Physiol Respir Environ
Exerc Physiol 57: 18571864, 1984. doi:10.1152/jappl.1984.57.6.1857.
93. Spiering BA, Mujika I, Sharp MA, Foulis SA.Maintaining physical
performance: the minimal dose of exercise needed to preserve
endurance and strength over time. JStrengthCondRes35: 1449
1458, 2021. doi:10.1519/JSC.0000000000003964.
94. McMaster DT, Gill N, Cronin J, McGuigan M.The development,
retention and decay rates of strength and power in elite rugby
union, rugby league and American football: a systematic review.
Sports Med 43: 367384, 2013. doi:10.1007/s40279-013-0031-3.
95. Mujika I, Padilla S.Muscular characteristics of detraining in humans.
Med Sci Sports Exerc 33: 12971303, 2001. doi:10.1097/00005768-
200108000-00009.
96. Mujika I.Intense training: the key to optimal performance before
and during the taper. Scand J Med Sci Sports 20: 2431, 2010.
doi:10.1111/j.1600-0838.2010.01189.x.
97. Mujika I, Sharma AP, Stellingwerff T.Contemporary periodization of
altitude training for elite endurance athletes: a narrative review.
Sports Med 49: 16511669, 2019. doi:10.1007/s40279-019-01165-y.
98. Gibala MJ.Physiological basis of interval training for performance
enhancement. Exp Physiol 106: 23242327, 2021. doi:10.1113/
EP088190.
99. Gibala MJ, Hawley JA.Sprinting toward tness. Cell Metab 25:
988990, 2017. doi:10.1016/j.cmet.2017.04.030.
100. MacInnis MJ, Gibala MJ.Physiological adaptations to interval train-
ing and the role of exercise intensity. J Physiol 595: 29152930,
2017. doi:10.1113/JP273196.
101. Burgomaster KA, Howarth KR, Phillips SM, Rakobowchuk M,
Macdonald MJ, McGee SL, Gibala MJ.Similar metabolic adapta-
tions during exercise after low volume sprint interval and traditional
endurance training in humans. JPhysiol586: 151160, 2008.
doi:10.1113/jphysiol.2007.142109.
102. Gibala MJ, Little JP, van Essen M, Wilkin GP, Burgomaster KA,
Safdar A, Raha S, Tarnopolsky MA.Short-term sprint interval versus
traditional endurance training: similar initial adaptations in human
skeletal muscle and exercise performance. J Physiol 575: 901911,
2006. doi:10.1113/jphysiol.2006.112094.
103. Lydiard A, Gilmour G. Run to the Top (2nd rev. ed.). London: Arena
Publications, 1968.
104. MacInnis MJ, Skelly LE, Gibala MJ.CrossTalk proposal: Exercise
training intensity is more important than volume to promote
increases in human skeletal muscle mitochondrial content. J
Physiol 597: 41114113, 2019. doi:10.1113/JP277633.
105. Bishop DJ, Botella J, Granata C.CrossTalk opposing view: Exercise
training volume is more important than training intensity to promote
increases in mitochondrial content. J Physiol 597: 41154118, 2019.
doi:10.1113/JP277634.
106. Flockhart M, Nilsson LC, Tais S, Ekblom B, Apr
oW,LarsenFJ.
Excessive exercise training causes mitochondrial functional impair-
ment and decreases glucose tolerance in healthy volunteers. Cell
Metab 33: 957970.e6, 2021. doi:10.1016/j.cmet.2021.02.017.
107. Hawley JA, Bishop DJ.High-intensity exercise training - too
much of a good thing? Nat Rev Endocrinol 17: 385386, 2021.
doi:10.1038/s41574-021-00500-6.
108. Matwejew LP.Die periodisierung des sportlichen trainings.
Leistungssport 2: 401409, 1972.
109. Mujika I, Halson S, Burke LM, Balagu
eG,FarrowD.An integrated,
multifactorial approach to periodization for optimal performance in
individual and team sports. Int J Sports Physiol Perform 13: 538
561, 2018. doi:10.1123/ijspp.2018-0093.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1761
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
110. Seiler S.What is best practice for training intensity and duration dis-
tribution in endurance athletes? Int J Sports Physiol Perform 5:
276291, 2010. doi:10.1123/ijspp.5.3.276.
111. Hawley JA, Burke LM. Peak Performance: Training and Nutritional
Strategies for Sport. St. Leonards, NSW, Australia: Allen & Unwin,
1998, p. 456.
112. Jones AM.Ave year physiological case study of an Olympic run-
ner. Br J Sports Med 32: 3943, 1998. doi:10.1136/bjsm.32.1.39.
113. Solli GS, Tønnessen E, Sandbakk O.The training characteristics of
the worlds most successful female cross-country skier. Front
Physiol 8: 1069, 2017. doi:10.3389/fphys.2017.01069.
114. Leo P, Spragg J, Simon D, Lawley JS, Mujika I.Training characteris-
tics and power prole of professional U23 cyclists throughout a
competitive season. Sports (Basel) 8: 120167, 2020. doi:10.3390/
sports8120167.
115. Casado A, González-Mohíno F, González-Rav
e JM, Foster C.
Training periodization, methods, intensity distribution, and volume
in highly trained and elite distance runners: a systematic review. Int
J Sports Physiol Perform 17: 820833, 2022. doi:10.1123/ijspp.
2021-0435.
116. Burnley M, Bearden SE, Jones AM.Polarized training is not opti-
mal for endurance athletes: response to foster and colleagues.
Med Sci Sports Exerc 54: 10381040, 2022. doi:10.1249/MSS.
0000000000002924.
117. Bourgois JG, Bourgois G, Boone J.Perspectives and determinants
for training-intensity distribution in elite endurance athletes. Int J
Sports Physiol Perform 14: 11511156, 2019. doi:10.1123/ijspp.2018-
0722.
118. Casado A, Hanley B, Ruiz-Perez LM.Deliberate practice in training
differentiates the best Kenyan and Spanish long-distance runners.
Eur J Sport Sci 20: 887895, 2020. doi:10.1080/17461391.2019.
1694077.
119. Casado A, Hanley B, Santos-Concejero J, Ruiz-Perez LM.World-
class long-distance running performances are best predicted by
volume of easy runs and deliberate practice of short-interval and
tempo runs. JStrengthCondRes35: 25252531, 2021. doi:10.1519/
JSC.0000000000003176.
120. Metcalfe AJ, Menaspà P, Villerius V, Quod M, Peiffer JJ, Govus AD,
Abbiss CR.Within-season distribution of external training and rac-
ing workload in professional male road cyclists. Int J Sports Physiol
Perform 12: S2142S2146, 2017. doi:10.1123/ijspp.2016-0396.
121. Seiler KS, Kjerland GO.Quantifying training intensity distribution in
elite endurance athletes: is there evidence for an optimaldistribu-
tion? Scand J Med Sci Sports 16: 4956, 2006. doi:10.1111/j.1600-
0838.2004.00418.x.
122. Billat V, Lepretre PM, Heugas AM, Laurence MH, Salim D,
Koralsztein JP.Training and bioenergetic characteristics in elite
male and female Kenyan runners. Med Sci Sports Exerc 35: 297
304, 2003. doi:10.1249/01.MSS.0000053556.59992.A9.
123. Billat VL, Demarle A, Slawinski J, Paiva M, Koralsztein JP.Physical
and training characteristics of top-class marathon runners. Med
Sci Sports Exerc 33: 20892097, 2001. doi:10.1097/00005768-
200112000-00018.
124. Karp JR.Training characteristics of qualiers for the U.S. Olympic
Marathon Trials. Int J Sports Physiol Perform 2: 7292, 2007.
doi:10.1123/ijspp.2.1.72.
125. Jeukendrup AE, Craig NP, Hawley JA.The bioenergetics of world
class cycling. JSciMedSport3: 414433, 2000. doi:10.1016/s1440-
2440(00)80008-0.
126. Sanders D, van Erp T, de Koning JJ.Intensity and load characteris-
tics of professional road cycling: differences between mensand
womensraces.Int J Sports Physiol Perform 14: 296302, 2019.
doi:10.1123/ijspp.2018-0190.
127. Mujika I, Chatard JC, Busso T, Geyssant A, Barale F, Lacoste L.
Effects of training on performance in competitive swimming. Can J
Appl Physiol 20: 395406, 1995. doi:10.1139/h95-031.
128. Rosenblat MA, Perrotta AS, Vicenzino B.Polarized vs. threshold
training intensity distribution on endurance sport performance: a
systematic review and meta-analysis of randomized controlled tri-
als. J Strength Cond Res 33: 34913500, 2019. doi:10.1519/
JSC.0000000000002618.
129. Coetzer P, Noakes TD, Sanders B, Lambert MI, Bosch AN, Wiggins
T, Dennis SC.Superior fatigue resistance of elite black South
African distance runners. J Appl Physiol (1985) 75: 18221827,
1993. doi:10.1152/jappl.1993.75.4.1822.
130. Guellich A, Seiler S, Emrich E.Training methods and intensity distri-
bution of young world-class rowers. Int J Sports Physiol Perform 4:
448460, 2009. doi:10.1123/ijspp.4.4.448.
131. Laursen PB.Training for intense exercise performance: high-inten-
sity or high-volume training? Scand J Med Sci Sports 20, Suppl 2:
110, 2010. doi:10.1111/j.1600-0838.2010.01184.x.
132. Foster C, Casado A, Esteve-Lanao J, Haugen T, Seiler S.Polarized
training is optimal for endurance athletes. Med Sci Sports Exerc 54:
10281031, 2022. doi:10.1249/MSS.0000000000002871.
133. Burnley M, Bearden SE, Jones AM.Polarized training is not optimal
for endurance athletes. Med Sci Sports Exerc 54: 10321034,
2022. doi:10.1249/MSS.0000000000002869.
134. Foster C, Casado A, Esteve-Lanao J, Haugen T, Seiler S.Polarized
training is optimal for endurance athletes: response to Burnley,
Bearden, and Jones. Med Sci Sports Exerc 54: 10351037, 2022.
doi:10.1249/MSS.0000000000002923.
135. Moesgaard L, Beck MM, Christiansen L, Aagaard P, Lundbye-
Jensen J.Effects of periodization on strength and muscle hypertro-
phy in volume-equated resistance training programs: a systematic
review and meta-analysis. Sports Med 52: 16471666, 2022.
doi:10.1007/s40279-021-01636-1.
136. Phillips SM.A brief review of critical processes in exercise-induced
muscular hypertrophy. Sports Med 44, Suppl 1: S71S77, 2014.
doi:10.1007/s40279-014-0152-3.
137. Lyristakis PM, Wundersitz DW, Zadow EK, Mnatzaganian G, Gordon
BA.The inuence of considering individual resistance training vari-
ables as a whole on muscle strength: a systematic review and
meta-analysis protocol. PLoS One 17: e0262674, 2022. doi:10.1371/
journal.pone.0262674.
138. Viecelli C, Aguayo D.May the force and mass be with youevi-
dence-based contribution of mechano-biological descriptors of re-
sistance exercise. Front Physiol 12: 686119, 2021. doi:10.3389/
fphys.2021.686119.
139. Schoenfeld BJ, Grgic J, Van Every DW, Plotkin DL.Loading recom-
mendations for muscle strength, hypertrophy, and local endurance:
a re-examination of the repetition continuum. Sports (Basel) 9: 32,
2021. doi:10.3390/sports9020032.
FURRER ET AL.
1762 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
140. American College of Sports Medicine. American College of Sports
Medicine position stand. Progression models in resistance training
for healthy adults. Med Sci Sports Exerc 41: 687708, 2009.
doi:10.1249/MSS.0b013e3181915670.
141. Alkner BA, Tesch PA, Berg HE.Quadriceps EMG/force relationship
in knee extension and leg press. Med Sci Sports Exerc 32: 459
463, 2000. doi:10.1097/00005768-200002000-00030.
142. Kumar V, Selby A, Rankin D, Patel R, Atherton P, Hildebrandt W,
Williams J, Smith K, Seynnes O, Hiscock N, Rennie MJ.Age-related
differences in the dose-response relationship of muscle protein
synthesis to resistance exercise in young and old men. J Physiol
587: 211217, 2009. doi:10.1113/jphysiol.2008.164483.
143. Burd NA, West DW, Staples AW, Atherton PJ, Baker JM, Moore DR,
Holwerda AM, Parise G, Rennie MJ, Baker SK, Phillips SM.Low-load
high volume resistance exercise stimulates muscle protein synthe-
sis more than high-load low volume resistance exercise in young
men. PLoS One 5: e12033, 2010. doi:10.1371/journal.pone.0012033.
144. Mitchell CJ, Churchward-Venne TA, West DW, Burd NA, Breen L,
Baker SK, Phillips SM.Resistance exercise load does not determine
training-mediated hypertrophic gains in young men. J Appl Physiol
(1985) 113: 7177, 2012. doi:10.1152/japplphysiol.00307.2012.
145. Alway SE, Sale DG, MacDougall JD.Twitch contractile adaptations
are not dependent on the intensity of isometric exercise in the
human triceps surae. Eur J Appl Physiol Occup Physiol 60: 346
352, 1990. doi:10.1007/BF00713497.
146. Haugen T, Sandbakk O, Enoksen E, Seiler S, nnessen E.
Crossing the golden training divide: the science and practice of
training world-class 800- and 1500-m runners. Sports Med 51:
18351854, 2021. doi:10.1007/s40279-021-01481-2.
147. Włodarczyk M, Adamus P, Zieli
nski J, Kantanista A.Effects of veloc-
ity-based training on strength and power in elite athletesasys-
tematic review. Int J Environ Res Public Health 18: 5257, 2021.
doi:10.3390/ijerph18105257.
148. Cormie P, McGuigan MR, Newton RU.Developing maximal neuro-
muscular power: part 2training considerations for improving maxi-
mal power production. Sports Med 41: 125146, 2011. doi:10.2165/
11538500-000000000-00000.
149. Patterson SD, Hughes L, Warmington S, Burr J, Scott BR, Owens J,
Abe T, Nielsen JL, Libardi CA, Laurentino G, Neto GR, Brandner C,
Martin-Hernandez J, Loenneke J.Blood Flow Restriction Exercise:
Considerations of Methodology, Application, and Safety. Front
Physiol 10: 533, 2019. doi:10.3389/fphys.2019.00533.
150. Pignanelli C, Christiansen D, Burr JF.Blood ow restriction training
and the high-performance athlete: science to application. JAppl
Physiol (1985) 130: 11631170, 2021. doi:10.1152/japplphysiol.00982.
2020.
151. Wortman RJ, Brown SM, Savage-Elliott I, Finley ZJ, Mulcahey
MK.Blood ow restriction training for athletes: a systematic
review. Am J Sports Med 49: 19381944, 2021. doi:10.1177/
0363546520964454.
152. Burd NA, Holwerda AM, Selby KC, West DW, Staples AW, Cain NE,
Cashaback JG, Potvin JR, Baker SK, Phillips SM.Resistance exer-
cise volume affects myobrillar protein synthesis and anabolic sig-
nalling molecule phosphorylation in young men. J Physiol 588:
31193130, 2010. doi:10.1113/jphysiol.2010.192856.
153. Williams TD, Tolusso DV, Fedewa MV, Esco MR.Comparison of
periodized and non-periodized resistance training on maximal
strength: a meta-analysis. Sports Med 47: 20832100, 2017.
doi:10.1007/s40279-017-0734-y.
154. van der Zwaard S, Brocherie F, Jaspers RT.Under the hood: skele-
tal muscle determinants of endurance performance. Front Sports
Act Living 3: 719434, 2021. doi:10.3389/fspor.2021.719434.
155. van Wessel T, de Haan A, van der Laarse WJ, JaspersRT.The mus-
cle ber type-ber size paradox: hypertrophy or oxidative metabo-
lism? Eur J Appl Physiol 110: 665694, 2010. doi:10.1007/s00421-
010-1545-0.
156. Hickson RC.Interference of strength development by simultane-
ously training for strength and endurance. Eur J Appl Physiol
Occup Physiol 45: 255263, 1980. doi:10.1007/BF00421333.
157. Rønnestad BR, Mujika I.Optimizing strength training for running
and cycling endurance performance: a review. Scand J Med Sci
Sports 24: 603612, 2014. doi:10.1111/sms.12104.
158. Berryman N, Mujika I, Bosquet L.Concurrent training for sports per-
formance: the 2 sides of the medal. Int J Sports Physiol Perform
14: 279285, 2019. doi:10.1123/ijspp.2018-0103.
159. Bassel-Duby R, Olson EN.Signaling pathways in skeletal muscle
remodeling. Annu Rev Biochem 75: 1937, 2006. doi:10.1146/
annurev.biochem.75.103004.142622.
160. Egan B, Hawley JA, Zierath JR.SnapShot: Exercise metabolism.
Cell Metab 24: 342342.e1, 2016. doi:10.1016/j.cmet.2016.07.013.
161. Hoffman NJ, Parker BL, Chaudhuri R, Fisher-Wellman KH, Kleinert
M, Humphrey SJ, Yang P, Holliday M, Trefely S, Fazakerley DJ,
St
ockli J, Burcheld JG, Jensen TE, Jothi R, Kiens B, Wojtaszewski
JF, Richter EA, James DE.Global phosphoproteomic analysis of
human skeletal muscle reveals a network of exercise-regulated ki-
nases and AMPK substrates. Cell Metab 22: 922935, 2015.
doi:10.1016/j.cmet.2015.09.001.
162. Hood DA, Irrcher I, Ljubicic V, Joseph AM.Coordination of meta-
bolic plasticity in skeletal muscle. J Exp Biol 209: 22652275,
2006. doi:10.1242/jeb.02182.
163. Hoppeler H.Molecular networks in skeletal muscle plasticity. JExp
Biol 219: 205213, 2016. doi:10.1242/jeb.128207.
164. Perry CG, Hawley JA.Molecular basis of exercise-induced skeletal
muscle mitochondrial biogenesis: historical advances, current
knowledge, and future challenges. Cold Spring Harb Perspect
Med 8: a029686, 2018. doi:10.1101/cshperspect.a029686.
165. Yang Y, Creer A, Jemiolo B, Trappe S.Time course of myogenic
and metabolic gene expression in response to acute exercise in
human skeletal muscle. J Appl Physiol (1985) 98: 17451752,
2005. doi:10.1152/japplphysiol.01185.2004.
166. Hughes DC, Ellefsen S, Baar K.Adaptations to endurance and
strength training. Cold Spring Harb Perspect Med 8: a029769,
2018. doi:10.1101/cshperspect.a029769.
167. Lundberg TR, Feuerbacher JF, S
unkeler M, Schumann M.The
effects of concurrent aerobic and strength training on muscle ber
hypertrophy: a systematic review and meta-analysis. Sports Med
52: 23912403, 2022. doi:10.1007/s40279-022-01688-x.
168. van der Zwaard S, Weide G, Levels K, Eikelboom MRI, Noordhof
DA, Hofmijster MJ, van der Laarse WJ, de Koning JJ, de Ruiter CJ,
Jaspers RT.Muscle morphology of the vastus lateralis is strongly
related to ergometer performance, sprint capacity and endurance
capacity in Olympic rowers. JSportsSci36: 21112120, 2018.
doi:10.1080/02640414.2018.1439434.
169. van der Zwaard S, van der Laarse WJ, Weide G, Bloemers FW,
Hofmijster MJ, Levels K, Noordhof DA, de Koning JJ, De Ruiter
CJ, Jaspers RT.Critical determinants of combined sprint and
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1763
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
endurance performance: an integrative analysis from muscle
ber to the human body. FASEB J 32: 21102123, 2018.
doi:10.1096/fj.201700827R.
170. Lundby C, Millet GP, Calbet JA, B
artsch P, Subudhi AW.Does alti-
tude trainingincrease exercise performance in elite athletes? Br J
Sports Med 46: 792795, 2012. doi:10.1136/bjsports-2012-091231.
171. Lundby C, Robach P.Does altitude trainingincrease exercise
performance in elite athletes? Exp Physiol 101: 783788, 2016.
doi:10.1113/EP085579.
172. Gore CJ, Sharpe K, Garvican-Lewis LA, Saunders PU, Humberstone
CE, Robertson EY, Wachsmuth NB, Clark SA, McLean BD,
Friedmann-Bette B, Neya M, Pottgiesser T, Schumacher YO,
Schmidt WF.Altitude training and haemoglobin mass from the opti-
mised carbon monoxide rebreathing method determined by a
meta-analysis. Br J Sports Med 47: i31i39, 2013. doi:10.1136/
bjsports-2013-092840.
173. Bergeron MF, Bahr R, B
artsch P, Bourdon L, Calbet JA, Carlsen KH,
Castagna O, González-Alonso J, Lundby C, Maughan RJ, Millet G,
Mountjoy M, Racinais S, Rasmussen P, Singh DG, Subudhi AW,
Young AJ, Soligard T, Engebretsen L.International Olympic
Committee consensus statement on thermoregulatory and altitude
challenges for high-level athletes. Br J Sports Med 46: 770779,
2012. doi:10.1136/bjsports-2012-091296.
174. Wehrlin JP, Marti B, Hall
en J.Hemoglobin mass and aerobic per-
formance at moderate altitude in elite athletes. Adv Exp Med Biol
903: 357374, 2016. doi:10.1007/978-1-4899-7678-9_24.
175. Levine BD, Stray-Gundersen J.Living high-training low: effect of
moderate-altitude acclimatization with low-altitude training on per-
formance. J Appl Physiol (1985) 83: 102112, 1997. doi:10.1152/
jappl.1997.83.1.102.
176. Saunders PU, Garvican-Lewis LA, Schmidt WF, Gore CJ.Relationship
between changes in haemoglobin mass and maximal oxygen uptake
after hypoxic exposure. Br J Sports Med 47, Suppl 1: i26i30, 2013.
doi:10.1136/bjsports-2013-092841.
177. Siebenmann C, Robach P, Jacobs RA, Rasmussen P, Nordsborg N,
Diaz V, Christ A, Olsen NV, Maggiorini M, Lundby C.Live high-train
lowusing normobaric hypoxia: a double-blinded, placebo-con-
trolled study. J Appl Physiol (1985) 112: 106117, 2012. doi:10.1152/
japplphysiol.00388.2011.
178. Bejder J, Andersen AB, Buchardt R, Larsson TH, Olsen NV,
Nordsborg NB.Endurance, aerobic high-intensity, and repeated
sprint cycling performance is unaffected by normobaric live high-
train low: a double-blind placebo-controlled cross-over study. Eur
J Appl Physiol 117: 979988, 2017. doi:10.1007/s00421-017-3586-0.
179. Gore CJ.The challenge of assessing athlete performance after alti-
tude training. J Appl Physiol (1985) 116: 593594, 2014. doi:10.1152/
japplphysiol.00029.2014.
180. Hurst P, Schipof-Godart L, Szabo A, Raglin J, Hettinga F, Roelands
B, Lane A, Foad A, Coleman D, Beedie C.The placebo and nocebo
effect on sports performance: a systematic review. Eur J Sport Sci
20: 279292, 2020. doi:10.1080/17461391.2019.1655098.
181. Chapman RF, Stickford JL, Levine BD.Altitude training considera-
tions for the winter sport athlete. Exp Physiol 95: 411421, 2010.
doi:10.1113/expphysiol.2009.050377.
182. Alvarez Villela M, Dunworth SA, Kraft BD, Harlan NP, Natoli MJ,
Suliman HB, Moon RE.Effects of high-intensity interval training with
hyperbaric oxygen. Front Physiol 13: 963799, 2022. doi:10.3389/
fphys.2022.963799.
183. van Doorslaer de Ten Ryen S, Warnier G, Gnimassou O, Belhaj MR,
Benoit N, Naslain D, Brook MS, Smith K, Wilkinson DJ, Nielens H,
Atherton PJ, Francaux M, Deldicque L.Higher strength gain after
hypoxic vs normoxic resistance training despite no changes in mus-
cle thickness and fractional protein synthetic rate. FASEB J 35:
e21773, 2021. doi:10.1096/fj.202100654RR.
184. Martin PM, Bart RM, Ashley RL, Velasco T, Wise SR.An overview
of blood ow restriction physiology and clinical considerations.
Curr Sports Med Rep 21: 123128, 2022. doi:10.1249/JSR.
0000000000000948.
185. OBrien L, Jacobs I.Methodological variations contributing to heter-
ogenous ergogenic responses to ischemic preconditioning. Front
Physiol 12: 656980, 2021. doi:10.3389/fphys.2021.656980.
186. Bartlett JD, Hawley JA, Morton JP.Carbohydrate availability and
exercise training adaptation: too much of a good thing? Eur J Sport
Sci 15: 312, 2015. doi:10.1080/17461391.2014.920926.
187. Hansen AK, Fischer CP, Plomgaard P, Andersen JL, Saltin B,
Pedersen BK.Skeletal muscle adaptation: training twice every sec-
ond day vs. training once daily. J Appl Physiol (1985) 98: 9399,
2005. doi:10.1152/japplphysiol.00163.2004.
188. Hawley JA, Burke LM.Carbohydrate availability and training adap-
tation: effects on cell metabolism. Exerc Sport Sci Rev 38: 152160,
2010. doi:10.1097/JES.0b013e3181f44dd9.
189. Hawley JA, Morton JP.Ramping up the signal: promoting endur-
ance training adaptation in skeletal muscle by nutritional manipula-
tion. Clin Exp Pharmacol Physiol 41: 608613, 2014. doi:10.1111/
1440-1681.12246.
190. Lane SC, Camera DM, Lassiter DG, Areta JL, Bird SR, Yeo WK,
Jeacocke NA, Krook A, Zierath JR, Burke LM, Hawley JA.Effects of
sleeping with reduced carbohydrate availability on acute training
responses. J Appl Physiol (1985) 119: 643655, 2015. doi:10.1152/
japplphysiol.00857.2014.
191. Psilander N, Frank P, Flockhart M, Sahlin K.Exercise with low glyco-
gen increases PGC-1alpha gene expression in human skeletal mus-
cle. Eur J Appl Physiol 113: 951963, 2013. doi:10.1007/s00421-012-
2504-8.
192. Wojtaszewski JF, Jørgensen SB, Frosig C, MacDonald C, Birk JB,
Richter EA.Insulin signalling: effects of prior exercise. Acta Physiol
Scand 178: 321328, 2003. doi:10.1046/j.1365-201X.2003.01151.x.
193. Edinburgh RM, Koumanov F, Gonzalez JT.Impact of pre-exercise
feeding status on metabolic adaptations to endurance-type exercise
training. JPhysiol600: 13271338, 2022. doi:10.1113/JP280748.
194. Hawley JA, Leckey JJ.Carbohydrate dependence during pro-
longed, intense endurance exercise. Sports Med 45, Suppl 1: S5
S12, 2015. doi:10.1007/s40279-015-0400-1.
195. Burke LM, Hawley JA, Wong SH, Jeukendrup AE.Carbohydrates
for training and competition. JSportsSci29, Suppl 1: S17S27, 2011.
doi:10.1080/02640414.2011.585473.
196. Burke LM, Hawley JA, Jeukendrup A, Morton JP, Stellingwerff T,
Maughan RJ.Toward a common understanding of diet-exercise
strategies to manipulate fuel availability for training and competition
preparation in endurance sport. Int J Sport Nutr Exerc Metab 28:
451463, 2018. doi:10.1123/ijsnem.2018-0289.
197. Marquet LA, Brisswalter J, Louis J, Tiollier E, Burke LM, Hawley JA,
Hausswirth C.Enhanced endurance performance by periodization
of carbohydrate intake: sleep lowstrategy. Med Sci Sports Exerc
48: 663672, 2016. doi:10.1249/MSS.0000000000000823.
FURRER ET AL.
1764 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
198. Marquet LA, Hausswirth C, Molle O, Hawley JA, Burke LM, Tiollier E,
Brisswalter J.Periodization of carbohydrate intake: short-term
effect on performance. Nutrients 8: 755, 2016. doi:10.3390/
nu8120755.
199. Yeo WK, McGee SL, Carey AL, Paton CD, Garnham AP, Hargreaves
M, Hawley JA.Acute signalling responses to intense endurance
training commenced with low or normal muscle glycogen. Exp
Physiol 95: 351358, 2010. doi:10.1113/expphysiol.2009.049353.
200. Hulston CJ, Venables MC, Mann CH, Martin C, Philp A, Baar K,
Jeukendrup AE.Training with low muscle glycogen enhances fat
metabolism in well-trained cyclists. Med Sci Sports Exerc 42:
20462055, 2010. doi:10.1249/MSS.0b013e3181dd5070.
201. Yeo WK, Paton CD, Garnham AP, Burke LM, Carey AL, Hawley JA.
Skeletal muscle adaptation and performance responses to once a
day versus twice every second day endurance training regimens. J
Appl Physiol (1985) 105: 14621470, 2008. doi:10.1152/japplphysiol.
90882.2008.
202. Burke LM, Ross ML, Garvican-Lewis LA, Welvaert M, Heikura IA,
Forbes SG, Mirtschin JG, Cato LE, Strobel N, Sharma AP, Hawley JA.
Low carbohydrate, high fat diet impairs exercise economy and neg-
ates the performance benetfromintensied training in elite race
walkers. JPhysiol595: 27852807, 2017. doi:10.1113/JP273230.
203. Burke LM, Sharma AP, Heikura IA, Forbes SF, Holloway M, McKay
AK, Bone JL, Leckey JJ, Welvaert M, Ross ML.Crisis of condence
averted: impairment of exercise economy and performance in elite
race walkers by ketogenic low carbohydrate, high fat (LCHF) diet is
reproducible. PLoS One 15: e0234027, 2020. doi:10.1371/journal.
pone.0234027.
204. Riis S, Møller AB, Dollerup O, Høffner L, Jessen N, Madsen K.Acute
and sustained effects of a periodized carbohydrate intake using the
sleep-low model in endurance-trained males. Scand J Med Sci
Sports 29: 18661880, 2019. doi:10.1111/sms.13541.
205. Gejl KD, Nybo L.Performance effects of periodized carbohydrate
restriction in endurance trained athletesasystematicreviewand
meta-analysis. J Int Soc Sports Nutr 18: 37, 2021. doi:10.1186/
s12970-021-00435-3.
206. Camera DM, Hawley JA, Coffey VG.Resistance exercise with low
glycogen increases p53 phosphorylation and PGC-1alpha mRNA in
skeletal muscle. Eur J Appl Physiol 115: 11851194, 2015. doi:10.1007/
s00421-015-3116-x.
207. Camera DM, West DW, Burd NA, Phillips SM, Garnham AP, Hawley
JA, Coffey VG.Low muscle glycogen concentration does not sup-
press the anabolic response to resistance exercise. J Appl Physi ol
(1985) 113: 206214, 2012. doi:10.1152/japplphysiol.00395.2012.
208. Asher G, Sassone-Corsi P.Time for food: the intimate interplay
between nutrition, metabolism, and the circadian clock. Cell 161:
8492, 2015. doi:10.1016/j.cell.2015.03.015.
209. Dyar KA, Ciciliot S, Wright LE, Biensø RS, Tagliazucchi GM, Patel
VR, Forcato M, Paz MI, Gudiksen A, Solagna F, Albiero M, Moretti I,
Eckel-Mahan KL, Baldi P, Sassone-Corsi P, Rizzuto R, Bicciato S,
Pilegaard H, Blaauw B, Schiafno S.Muscle insulin sensitivity and
glucose metabolism are controlled by the intrinsic muscle clock.
Mol Metab 3: 2941, 2014. doi:10.1016/j.molmet.2013.10.005.
210. Dyar KA, Lutter D, Artati A, Ceglia NJ, Liu Y, Armenta D, Jastroch M,
Schneider S, de Mateo S, Cervantes M, Abbondante S, Tognini P,
Orozco-Solis R, Kinouchi K, Wang C, Swerdloff R, Nadeef S, Masri S,
Magistretti P, Orlando V, Borrelli E, Uhlenhaut NH, Baldi P, Adamski
J, Tsch
op MH, Eckel-Mahan K, Sassone-Corsi P.Atlas of circadian
metabolism reveals system-wide coordination and communication
between clocks. Cell 174: 15711585.e11, 2018. doi:10.1016/j.cell.2018.
08.042.
211. Smith JG, Sassone-Corsi P.Clock-in, clock-out: circadian timekeep-
ing between tissues. Biochem (Lond) 42: 610, 2020. doi:10.1042/
bio04202007.
212. Adamovich Y, Dandavate V, Ezagouri S, Manella G, Zwighaft Z,
Sobel J, Kuperman Y, Golik M, Auerbach A, Itkin M, Malitsky S, Asher
G.Clock proteins and training modify exercise capacity in a daytime-
dependent manner. Proc Natl Acad Sci USA 118: e2101115118, 2021.
doi:10.1073/pnas.2101115118.
213. Kemler D, Wolff CA, Esser KA.Time-of-day dependent effects of
contractile activity on the phase of the skeletal muscle clock. J
Physiol 598: 36313644, 2020. doi:10.1113/JP279779.
214. Sato S, Basse AL, Sch
onke M, Chen S, Samad M, AltıntasA, Laker
RC, Dalbram E, Barr
es R, Baldi P, Treebak JT, Zierath JR, Sassone-
Corsi P.Time of exercise species the impact on muscle metabolic
pathways and systemic energy homeostasis. Cell Metab 30: 92
110.e4, 2019. doi:10.1016/j.cmet.2019.03.013.
215. Wolff CA, Esser KA.Exercise timing and circadian rhythms. Curr
Opin Physiol 10: 6469, 2019. doi:10.1016/j.cophys.2019.04.020.
216. Wolff CA, Esser KA.Exercise sets the muscle clock with a calcium
assist. J Physiol 598: 55915592, 2020. doi:10.1113/JP280783.
217. Eckel-Mahan KL, Patel VR, de Mateo S, Orozco-Solis R, Ceglia NJ,
Sahar S, Dilag-Penilla SA, Dyar KA, Baldi P, Sassone-Corsi P.
Reprogramming of the circadian clock by nutritional challenge. Cell
155: 14641478, 2013. doi:10.1016/j.cell.2013.11.034.
218. Greco CM, Koronowski KB, Smith JG, Shi J, Kunderfranco P,
Carriero R, Chen S, Samad M, Welz PS, Zinna VM, Mortimer T, Chun
SK, Shimaji K, Sato T, Petrus P, Kumar A, Vaca-Dempere M,
Deryagin O, Van C, Kuhn JM, Lutter D, Seldin MM, Masri S, Li W,
Baldi P, Dyar KA, Mu~
noz-Cánoves P, Benitah SA, Sassone-Corsi P.
Integration of feeding behavior by the liver circadian clock reveals
network dependency of metabolic rhythms. Sci Adv 7: eabi7828,
2021. doi:10.1126/sciadv.abi7828.
219. Lundell LS, Parr EB, Devlin BL, Ingerslev LR, AltıntasA, Sato S,
Sassone-Corsi P, Barr
es R, Zierath JR, Hawley JA.Time-restricted
feeding alters lipid and amino acid metabolite rhythmicity without
perturbing clock gene expression. Nat Commun 11: 4643, 2020.
doi:10.1038/s41467-020-18412-w.
220. Sato S, Parr EB, Devlin BL, Hawley JA, Sassone-Corsi P.Human
metabolomics reveal daily variations under nutritional challenges
specic to serum and skeletal muscle. Mol Metab 16: 111, 2018.
doi:10.1016/j.molmet.2018.06.008.
221. Atkinson G, Reilly T.Circadian variation in sports performance.
Sports Med 21: 292312, 1996. doi:10.2165/00007256-199621040-
00005.
222. de Goede P, Wefers J, Brombacher EC, Schrauwen P, Kalsbeek A.
Circadian rhythms in mitochondrial respiration. J Mol Endocrinol
60: R115R130, 2018. doi:10.1530/JME-17-0196.
223. Drust B, Waterhouse J, Atkinson G, Edwards B, Reilly T.Circadian
rhythms in sports performancean update. Chronobiol Int 22: 21
44, 2005. doi:10.1081/cbi-200041039.
224. Facer-Childs E, Brandstaetter R.The impact of circadian phenotype
and time since awakening on diurnal performance in athletes. Curr
Biol 25: 518522, 2015. doi:10.1016/j.cub.2014.12.036.
225. Kusumoto H, Ta C, Brown SM, Mulcahey MK.Factors contribut-
ing to diurnal variation in athletic performance and methods to
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1765
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
reduce within-day performance variation: a systematic review.
JStrengthCondRes35: S119S135, 2021. doi:10.1519/JSC.
0000000000003758.
226. Knaier R, Infanger D, Cajochen C, Schmidt-Trucksaess A, Faude O,
Roth R.Diurnal and day-to-day variations in isometric and isokinetic
strength. Chronobiol Int 36: 15371549, 2019. doi:10.1080/07420528.
2019.1658596.
227. Knaier R, Infanger D, Niemeyer M, Cajochen C, Schmidt-Trucks
ass
A.In athletes, the diurnal variations in maximum oxygen uptake are
more than twice as large as the day-to-day variations. Front Physiol
10: 219, 2019. doi:10.3389/fphys.2019.00219.
228. Roenneberg T, Wirz-Justice A, Merrow M.Life between clocks: daily
temporal patterns of human chronotypes. JBiolRhythms18: 80
90, 2003. doi:10.1177/0748730402239679.
229. Bailey SL, Heitkemper MM.Circadian rhythmicity of cortisol and
body temperature: morningness-eveningness effects. Chronobiol
Int 18: 249261, 2001. doi:10.1081/cbi-100103189.
230. Takahashi JS, Hong HK, Ko CH, McDearmon EL.The genetics of
mammalian circadian order and disorder: implications for physiol-
ogy and disease. Nat Rev Genet 9: 764775, 2008. doi:10.1038/
nrg2430.
231. Facer-Childs ER, Boiling S, Balanos GM.The effects of time of day
and chronotype on cognitive and physical performance in healthy
volunteers. Sports Med Open 4: 47, 2018. doi:10.1186/s40798-018-
0162-z.
232. Ammar A, Chtourou H, Souissi N.Effect of time-of-day on biochemi-
cal markers in response to physical exercise. JStrengthCondRes
31: 272282, 2017. doi:10.1519/JSC.0000000000001481.
233. Chtourou H, Driss T, Souissi S, Gam A, Chaouachi A, Souissi N.The
effect of strength trainingat the same time of the day on the diurnal
uctuations of muscular anaerobic performances. JStrengthCond
Res 26: 217225, 2012. doi:10.1519/JSC.0b013e31821d5e8d.
234. Chtourou H, Souissi N.The effect of training at a specictimeof
day: a review. J Strength Cond Res 26: 19842005, 2012.
doi:10.1519/JSC.0b013e31825770a7.
235. K
u
usmaa M, Schumann M, Sedliak M, Kraemer WJ, Newton RU,
Malinen JP, Nyman K, H
akkinen A, H
akkinen K.Effects of morning
versus evening combined strength and endurance training on
physical performance, muscle hypertrophy, and serum hormone
concentrations. Appl Physiol Nutr Metab 41: 12851294, 2016.
doi:10.1139/apnm-2016-0271.
236. Knaier R, Qian J, Roth R, Infanger D, Notter T, Wang W, Cajochen C,
Scheer F.Diurnal variation in maximum endurance and maximum
strength performance: a systematic review and meta-analysis.
Med Sci Sports Exerc 54: 169180, 2022. doi:10.1249/MSS.
0000000000002773.
237. van Moorsel D, Hansen J, Havekes B, Scheer F, J
orgensen JA,
Hoeks J, Schrauwen-Hinderling VB, Duez H, Lefebvre P, Schaper
NC, Hesselink MK, Staels B, Schrauwen P.Demonstration of a day-
night rhythm in human skeletal muscle oxidative capacity. Mol
Metab 5: 635645, 2016. doi:10.1016/j.molmet.2016.06.012.
238. Lok R, Zerbini G, Gordijn MC, Beersma DG, Hut RA.Gold, silver or
bronze: circadian variation strongly affects performance in Olympic
athletes. Sci Rep 10: 16088, 2020. doi:10.1038/s41598-020-72573-8.
239. Haugen T, Sandbakk O, Seiler S, nnessen E.The training charac-
teristics of world-class distance runners: an integration of scientic
literature and results-proven practice. Sports Med Open 8: 46,
2022. doi:10.1186/s40798-022-00438-7.
240. Haugen T, Seiler S, Sandbakk O, Tønnessen E.The training and de-
velopment of elite sprint performance: an integration of scientic
and best practice literature. Sports Med Open 5: 44, 2019.
doi:10.1186/s40798-019-0221-0.
241. Ansdell P, Thomas K, Hicks KM, Hunter SK, Howatson G, Goodall S.
Physiological sex differences affect the integrative response to
exercise: acute and chronic implications. Exp Physiol 105: 2007
2021, 2020. doi:10.1113/EP088548.
242. Landen S, Voisin S, Craig JM, McGee SL, Lamon S, Eynon N.
Genetic and epigenetic sex-specic adaptations to endurance
exercise. Epigenetics 14: 523535, 2019. doi:10.1080/15592294.
2019.1603961.
243. Besson T, Macchi R, Rossi J, Morio CY, Kunimasa Y, Nicol C,
Vercruyssen F, Millet GY.Sex differences in endurance running.
Sports Med 52: 12351257, 2022. doi:10.1007/s40279-022-01651-w.
244. Sims ST, Heather AK.Myths and methodologies: reducing scientic
design ambiguity in studies comparing sexes and/or menstrual
cycle phases. Exp Physiol 103: 13091317, 2018. doi:10.1113/
EP086797.
245. Nuzzo JL.Narrative review of sex differences in muscle strength,
endurance, activation, size, ber type, and strength training partici-
pation rates, preferences, motivations, injuries, and neuromuscular
adaptations. JStrengthCondRes37: 494536, 2023. doi:10.1519/
JSC.0000000000004329.
246. Cullen ML, Casazza GA, Davis BA.Passive recovery strategies af-
ter exercise: a narrative literature review of the current evidence.
Curr Sports Med Rep 20: 351358, 2021. doi:10.1249/JSR.
0000000000000859.
247. Ihsan M, Abbiss CR, Allan R.Adaptations to post-exercise cold
water immersion: friend, foe, or futile? Front Sports Act Living 3:
714148, 2021. doi:10.3389/fspor.2021.714148.
248. Kissling LS, Akerman AP, Cotter JD.Heat-induced hypervolemia:
does the mode of acclimation matter and what are the implications
for performance at Tokyo 2020? Temperature (Austin) 7: 129148,
2019. doi:10.1080/23328940.2019.1653736.
249. Rønnestad BR, Hamarsland H, Hansen J, Holen E, Montero D,
Whist JE, Lundby C.Five weeks of heat training increases haemo-
globin mass in elite cyclists. Exp Physiol 106: 316327, 2021.
doi:10.1113/EP088544.
250. P
eriard JD, Eijsvogels TM, Daanen HA.Exercise under heat stress:
thermoregulation, hydration, performance implications, and mitiga-
tion strategies. Physiol Rev 101: 18731979, 2021. doi:10.1152/
physrev.00038.2020.
251. Joyner MJ, Casey DP.Regulation of increased blood ow (hypere-
mia) to muscles during exercise: a hierarchy of competing physio-
logical needs. Physiol Rev 95: 549601, 2015. doi:10.1152/
physrev.00035.2013.
252. Haugen T, Paulsen G, Seiler S, Sandbakk O.New records in human
power. Int J Sports Physiol Perform 13: 678686, 2018. doi:10.1123/
ijspp.2017-0441.
253. Zhou B, Conlee RK, Jensen R, Fellingham GW, George JD, Fisher
AG.Stroke volume does not plateau during graded exercise in elite
male distance runners. Med Sci Sports Exerc 33: 18491854, 2001.
doi:10.1097/00005768-200111000-00008.
254. Dandanell S, Meinild-Lundby AK, Andersen AB, Lang PF, Oberholzer
L, Keiser S, Robach P, Larsen S, Rønnestad BR, Lundby C.
Determinants of maximal whole-body fat oxidation in elite cross-
FURRER ET AL.
1766 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
country skiers: Role of skeletal muscle mitochondria. Scand J Med
Sci Sports 28: 24942504, 2018. doi:10.1111/sms.13298.
255. Hoppeler H, Howald H, Conley K, Lindstedt SL, Claassen H, Vock P,
Weibel ER.Endurance training in humans: aerobic capacity and
structure of skeletal muscle. J Appl Physiol (1985) 59: 320327,
1985. doi:10.1152/jappl.1985.59.2.320.
256. Ørtenblad N, Nielsen J, Saltin B, Holmberg HC.Role of glycogen
availability in sarcoplasmic reticulum Ca
21
kinetics in human skele-
tal muscle. J Physiol 589: 711725, 2011. doi:10.1113/jphysiol.2010.
195982.
257. Jones AM.The physiology of the world record holder for the
womensmarathon.Int J Sports Sci Coaching 1: 101116, 2006.
doi:10.1260/174795406777641258.
258. Sandbakk O, Holmberg HC.Physiological capacity and training rou-
tines of elite cross-country skiers: approaching the upper limits of
human endurance. Int J Sports Physiol Perform 12: 10031011,
2017. doi:10.1123/ijspp.2016-0749.
259. Tønnessen E, Sylta Ø, Haugen TA, Hem E, Svendsen IS, Seiler S.
The road to gold: training and peaking characteristics in the year
prior to a gold medal endurance performance. PLoS One 9:
e101796, 2014. doi:10.1371/journal.pone.0101796.
260. Prommer N, Thoma S, Quecke L, Gutekunst T, V
olzke C,
Wachsmuth N, Niess AM, Schmidt W.Total hemoglobin mass and
blood volume of elite Kenyan runners. Med Sci Sports Exerc 42:
791797, 2010. doi:10.1249/MSS.0b013e3181badd67.
261. Lucía A, Rabadán M, Hoyos J, Hernández-Capilla M, P
erez M, San
Juan AF, Earnest CP, Chicharro JL.Frequency of the Vo
2max
pla-
teau phenomenon in world-class cyclists. Int J Sports Med 27:
984992, 2006. doi:10.1055/s-2006-923833.
262. Pate RR, Sparling PB, Wilson GE, Cureton KJ, Miller BJ.Cardiore-
spiratory and metabolic responses to submaximal and maximal exer-
cise in elite women distance runners. Int J Sports Med 8, Suppl 2: 91
95, 1987. doi:10.1055/s-2008-1025712.
263. Daniels J, Krahenbuhl G, Foster C, Gilbert J, Daniels S.Aerobic
responses of female distance runners to submaximal and maximal
exercise. Ann NY Acad Sci 301: 726733, 1977. doi:10.1111/j.1749-
6632.1977.tb38242.x.
264. Mahler DA, Shuhart CR, Brew E, Stukel TA.Ventilatory responses
and entrainment of breathing during rowing. Med Sci Sports Exerc
23: 186192, 1991.
265. Wang E, Solli GS, Nyberg SK, Hoff J, Helgerud J.Stroke volume
does not plateau in female endurance athletes. Int J Sports Med
33: 734739, 2012. doi:10.1055/s-0031-1301315.
266. Degens H, Stasiulis A, Skurvydas A, Statkeviciene B, Venckunas T.
Physiological comparison between non-athletes, endurance, power
and team athletes. Eur J Appl Physiol 119: 13771386, 2019.
doi:10.1007/s00421-019-04128-3.
267. Burtscher M, Nachbauer W, Wilber R.The upper limit of aerobic
power in humans. Eur J Appl Physiol 111: 26252628, 2011.
doi:10.1007/s00421-011-1885-4.
268. Folinsbee LJ, Wallace ES, Bedi JF, Horvath SM.Exercise respiratory
pattern in elite cyclists and sedentary subjects. Med Sci Sports
Exerc 15: 503509, 1983.
269. Lucía A, Carvajal A, Calder
on FJ, Alfonso A, Chicharro JL.Breathing
pattern in highly competitive cyclists during incremental exercise.
Eur J Appl Physiol Occup Physiol 79: 512521, 1999. doi:10.1007/
s004210050546.
270. Lucia A, Hoyos J, Chicharro JL.Physiology of professional road cy-
cling. Sports Med 31: 325337, 2001. doi:10.2165/00007256-
200131050-00004.
271. Marinho AH, Cristina-Souza G, Santos PS, Santos-Mariano AC,
Rodacki A, De-Oliveira FR, Bertuzzi R, Lima-Silva AE.Caffeine alters
the breathing pattern during high-intensity whole-body exercise in
healthy men. Eur J Appl Physiol 122: 14971507, 2022. doi:10.1007/
s00421-022-04934-2.
272. Olfert IM, Balouch J, Kleinsasser A, Knapp A, Wagner H, Wagner
PD, Hopkins SR.Does gender affect human pulmonary gas
exchange during exercise? J Physiol 557: 529541, 2004.
doi:10.1113/jphysiol.2003.056887.
273. Edvardsen E, Hansen BH, Holme IM, Dyrstad SM, Anderssen SA.
Reference values for cardiorespiratory response and tness on the
treadmill in a 20- to 85-year-old population. Chest 144: 241248,
2013. doi:10.1378/chest.12-1458.
274. Ekblom B, Hermansen L.Cardiac output in athletes. J Appl Physiol
25: 619625, 1968. doi:10.1152/jappl.1968.25.5.619.
275. St Pierre SR, Peirlinck M, Kuhl E.Sex matters: a comprehensive
comparison of female and male hearts. Front Physiol 13: 831179,
2022. doi:10.3389/fphys.2022.831179.
276. PatelHN,MiyoshiT,AddetiaK,HenryMP,CitroR,DaimonM,etal.
Normal values of cardiac output and stroke volume according to
measurement technique, age, sex, and ethnicity: results of the World
Alliance of Societies of Echocardiography Study. JAmSoc
Echocardiogr 34: 10771085.e1, 2021. doi:10.1016/j.echo.2021.05.012.
277. Coyle EF.Improved muscular efciency displayed as Tour de
France champion matures. J Appl Physiol (1985) 98: 21912196,
2005. doi:10.1152/japplphysiol.00216.2005.
278. Tesch PA, Thorsson A, Kaiser P.Muscle capillary supply and ber
type characteristics in weight and power lifters. J Appl Physiol
Respir Environ Exerc Physiol 56: 3538, 1984. doi:10.1152/
jappl.1984.56.1.35.
279. Saltin B, Henriksson J, Nygaard E, Andersen P, Jansson E.Fiber
types and metabolic potentials of skeletal muscles in sedentary
man and endurance runners. Ann NY Acad Sci 301: 329, 1977.
doi:10.1111/j.1749-6632.1977.tb38182.x.
280. Tesch PA, Karlsson J.Muscle ber types and size in trained and
untrained muscles of elite athletes. J Appl Physiol (1985) 59: 1716
1720, 1985. doi:10.1152/jappl.1985.59.6.1716.
281. Costill DL, Daniels J, Evans W, Fink W, Krahenbuhl G, Saltin B.
Skeletal muscle enzymes and ber composition in male and female
track athletes. J Appl Physiol 40: 149154, 1976. doi:10.1152/
jappl.1976.40.2.149.
282. Larsen S, Nielsen J, Hansen CN, Nielsen LB, Wibrand F, Stride N,
Schroder HD, Boushel R, Helge JW, Dela F, Hey-Mogensen M.
Biomarkers of mitochondrial content in skeletal muscle of healthy
young human subjects. JPhysiol590: 33493360, 2012. doi:10.1113/
jphysiol.2012.230185.
283. Costill DL, Fink WJ, Flynn M, Kirwan J.Muscle ber composition
and enzyme activities in elite female distance runners. Int J Sports
Med 8, Suppl 2: 103106, 1987. doi:10.1055/s-2008-1025714.
284. Staron RS, Hagerman FC, Hikida RS, Murray TF, Hostler DP, Crill
MT, Ragg KE, Toma K.Fiber type composition of the vastus lateralis
muscle of young men and women. J Histochem Cytochem 48:
623629, 2000. doi:10.1177/002215540004800506.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1767
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
285. Hoppeler H, L
uthi P, Claassen H, Weibel ER, Howald H.The ultra-
structure of the normal human skeletal muscle. A morphometric
analysis on untrained men, women and well-trained orienteers.
Pugers Arch 344: 217232, 1973. doi:10.1007/BF00588462.
286. Ørtenblad N, Nielsen J, Boushel R, S
oderlund K, Saltin B, Holmberg
HC.The muscle ber proles, mitochondrial content, and enzyme
activities of the exceptionally well-trained arm and leg muscles of
elite cross-country skiers. Front Physiol 9: 1031, 2018. doi:10.3389/
fphys.2018.01031.
287. Hostler D, Schwirian CI, Campos G, Toma K, Crill MT, Hagerman
GR, Hagerman FC, Staron RS.Skeletal muscle adaptations in elastic
resistance-trained young men and women. Eur J Appl Physiol 86:
112118, 2001. doi:10.1007/s004210100495.
288. Cardinale DA, Horwath O, Elings-Knutsson J, Helge T, Godhe M,
Bermon S, Moberg M, Flockhart M, Larsen FJ, Hirschberg AL,
Ekblom B.Enhanced skeletal muscle oxidative capacity and capil-
lary-to-ber ratio following moderately increased testosterone ex-
posure in young healthy women. Front Physiol 11: 585490, 2020.
doi:10.3389/fphys.2020.585490.
289. Meinild Lundby AK, Jacobs RA, Gehrig S, de Leur J, Hauser M,
Bonne TC, Fluck D, Dandanell S, Kirk N, Kaech A, Ziegler U, Larsen
S, Lundby C.Exercise training increases skeletal muscle mitochon-
drial volume density by enlargement of existing mitochondria and
not de novo biogenesis. Acta Physiol (Oxf) 222: e12905, 2018.
doi:10.1111/apha.12905.
290. Nielsen J, Gejl KD, Hey-Mogensen M, Holmberg HC, Suetta C,
Krustrup P, Elemans CP, Ørtenblad N.Plasticity in mitochondrial cris-
tae density allows metabolic capacity modulation in human skeletal
muscle. JPhysiol595: 28392847, 2017. doi:10.1113/JP273040.
291. Ferretti G, Fagoni N, Taboni A, Vinetti G, di Prampero PE.Acentury
of exercise physiology: key concepts on coupling respiratory oxy-
gen ow to muscle energy demand during exercise. Eur J Appl
Physiol 122: 13171365, 2022. doi:10.1007/s00421-022-04901-x.
292. Rønnestad BR, Hansen J, Stensløkken L, Joyner MJ, Lundby C.
Case studies in physiology: temporal changes in determinants of
aerobic performance in individual going from alpine skier to world
junior champion time trial cyclist. J Appl Physiol (1985) 127: 306
311, 2019. doi:10.1152/japplphysiol.00798.2018.
293. Weibel ER, Hoppeler H.Exercise-induced maximal metabolic rate
scales with muscle aerobic capacity. J Exp Biol 208: 16351644,
2005. doi:10.1242/jeb.01548.
294. Sandbakk O, Solli GS, Holmberg HC.Sex differences in world-re-
cord performance: the inuence of sport discipline and competition
duration. Int J Sports Physiol Perform 13: 28, 2018. doi:10.1123/
ijspp.2017-0196.
295. Wagner PD.Determinants of maximal oxygen consumption. J
Muscle Res Cell Motil. In Press. doi:10.1007/s10974-022-09636-y.
296. McKenzie DC.Respiratory physiology: adaptations to high-level
exercise. Br J Sports Med 46: 381384, 2012. doi:10.1136/bjsports-
2011-090824.
297. Dauty M, Georges T, Le Blanc C, Louguet B, Menu P, Fouasson-
Chailloux A.Reference values of forced vital capacity and expira-
tory ow in high-level cyclists. Life (Basel) 11: 1293, 2021.
doi:10.3390/life11121293.
298. Segizbaeva MO, Aleksandrova NP.Respiratory muscle strength
and ventilatory function outcome: differences between trained ath-
letes and healthy untrained persons. Adv Exp Med Biol 1289: 89
97, 2021. doi:10.1007/5584_2020_554.
299. Calbet JA, Rådegran G, Boushel R, Søndergaard H, Saltin B,
Wagner PD.Effect of blood haemoglobin concentration on V
O2,max
and cardiovascular function in lowlanders acclimatised to 5260 m.
J Physiol 545: 715728, 2002. doi:10.1113/jphysiol.2002.029108.
300. Montero D, Cathomen A, Jacobs RA, Fl
uck D, de Leur J, Keiser S,
Bonne T, Kirk N, Lundby AK, Lundby C.Haematological rather than
skeletal muscle adaptations contribute to the increase in peak oxy-
gen uptake induced by moderate endurance training. J Physiol
593: 46774688, 2015. doi:10.1113/JP270250.
301. Skattebo O, Calbet JA, Rud B, Capelli C, Hall
en J.Contribution of
oxygen extraction fraction to maximal oxygen uptake in healthy
young men. Acta Physiol (Oxf) 230: e13486, 2020. doi:10.1111/
apha.13486.
302. R
osler K, Hoppeler H, Conley KE, Claassen H, Gehr P, Howald H.
Transfer effects in endurance exercise. Adaptations in trained and
untrained muscles. Eur J Appl Physiol Occup Physiol 54: 355
362, 1985. doi:10.1007/BF02337178.
303. Fry AC, Schilling BK, Staron RS, Hagerman FC, Hikida RS, Thrush
JT.Muscle ber characteristics and performance correlates of male
Olympic-style weightlifters. JStrengthCondRes17: 746754,
2003. doi:10.1519/1533-4287(2003)017<0746:mfcapc>2.0.co;2.
304. Storey A, Smith HK.Unique aspects of competitive weightlifting:
performance, training and physiology. Sports Med 42: 769790,
2012. doi:10.1007/BF03262294.
305. Zaras N, Stasinaki AN, Spiliopoulou P, Hadjicharalambous M, Terzis
G.Lean body mass, muscle architecture, and performance in well-
trained female weightlifters. Sports (Basel) 8: 67, 2020. doi:10.3390/
sports8050067.
306. Heckman CJ, Enoka RM.Physiology of the motor neuron and the
motor unit. In: Handbook of Clinical Neurophysiology,editedbyEisenA.
Amsterdam:Elsevier, 2004, p.119147.
307. Enoka RM, Pearson KG.The motor unit and muscle action. In:
Principles of Neural Science (5th ed.). New York: McGraw-Hill
Education, 2014.
308. Zou S, Pan BX.Post-synaptic specialization of the neuromuscular
junction: junctional folds formation, function, and disorders. Cell
Biosci 12: 93, 2022. doi:10.1186/s13578-022-00829-z.
309. Ruff RL.Endplate contributions to the safety factor for neuromus-
cular transmission. Muscle Nerve 44: 854861, 2011. doi:10.1002/
mus.22177.
310. Wood SJ, Slater CR.Safety factor at the neuromuscular junction.
Prog Neurobiol 64: 393429, 2001. doi:10.1016/s0301-0082(00)
00055-1.
311. Enoka RM, Duchateau J.Rate coding and the control of muscle
force. Cold Spring Harb Perspect Med 7: a029702, 2017. doi:10.1101/
cshperspect.a029702.
312. Neto WK, Ciena AP, Anaruma CA, de Souza RR, Gama EF.Effects
of exercise on neuromuscular junction components across age:
systematic review of animal experimental studies. BMC Res Notes
8: 713, 2015. doi:10.1186/s13104-015-1644-4.
313. Deschenes MR.Adaptations of the neuromuscular junction to exer-
cise training. Curr Opin Physiol 10: 1016, 2019. doi:10.1016/j.
cophys.2019.02.004.
314. Deschenes MR, Kressin KA, Garratt RN, Leathrum CM, Shaffrey EC.
Effects of exercise training on neuromuscular junction morphology
and pre- to post-synaptic coupling in young and aged rats.
FURRER ET AL.
1768 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
Neuroscience 316: 167177, 2016. doi:10.1016/j.neuroscience.2015.
12.004.
315. Arnold AS, Gill J, Christe M, Ruiz R, McGuirk S, St-Pierre J, Tabares
L, Handschin C.Morphological and functional remodelling of the
neuromuscular junction by skeletal muscle PGC-1alpha. Nat
Commun 5: 3569, 2014. doi:10.1038/ncomms4569.
316. Jones RA, Harrison C, Eaton SL, Llavero Hurtado M, Graham LC,
Alkhammash L, Oladiran OA, Gale A, Lamont DJ, Simpson H,
Simmen MW, Soeller C, Wishart TM, Gillingwater TH.Cellular and
molecular anatomy of the human neuromuscular junction. Cell Rep
21: 23482356, 2017. doi:10.1016/j.celrep.2017.11.008.
317. Duchateau J, Enoka RM.Distribution of motor unit properties across
human muscles. J Appl Physiol (1985) 132: 113, 2022. doi:10.1152/
japplphysiol.00290.2021.
318. Duchateau J, Semmler JG, Enoka RM.Training adaptations in the
behavior of human motor units. J Appl Physiol (1985) 101: 1766
1775, 2006. doi:10.1152/japplphysiol.00543.2006.
319. Hudson AL, Gandevia SC, Butler JE.A principle of neurome-
chanical matching for motor unit recruitment in human move-
ment. Exerc Sport Sci Rev 47: 157168, 2019. doi:10.1249/JES.
0000000000000191.
320. Hug F, Avrillon S, Ibá~
nez J, Farina D.Common synaptic input, syner-
gies and size principle: control of spinal motor neurons for move-
ment generation. J Physiol 601: 1120, 2023. doi:10.1113/JP283698.
321. Bawa PN, Jones KE, Stein RB.Assessment of size ordered recruit-
ment. Front Hum Neurosci 8: 532, 2014. doi:10.3389/fnhum.2014.
00532.
322. Marshall NJ, Glaser JI, Trautmann EM, Amematsro EA, Perkins SM,
Shadlen MN, Abbott LF, Cunningham JP, Churchland MM.Flexible
neural control of motor units. Nat Neurosci 25: 14921504, 2022.
doi:10.1038/s41593-022-01165-8.
323. Adam A, De Luca CJ.Firing rates of motor units in human vastus
lateralis muscle during fatiguing isometric contractions. J Appl
Physiol (1985) 99: 268280, 2005. doi:10.1152/japplphysiol.
01344.2004.
324. Morton RW, Sonne MW, Farias Zuniga A, Mohammad IY, Jones A,
McGlory C, Keir PJ, Potvin JR, Phillips SM.Muscle bre activation is
unaffected by load and repetition duration when resistance exer-
cise is performed to task failure. J Physiol 597: 46014613, 2019.
doi:10.1113/JP278056.
325. Morton RW, Oikawa SY, Wavell CG, Mazara N, McGlory C,
Quadrilatero J, Baechler BL, Baker SK, Phillips SM.Neither load nor
systemic hormones determine resistance training-mediated hyper-
trophy or strength gains in resistance-trained young men. JAppl
Physiol (1985) 121: 129138, 2016. doi:10.1152/japplphysiol.00154.
2016.
326. Blum JA, Klemm S, Shadrach JL, Guttenplan KA, Nakayama L,
Kathiria A, Hoang PT, Gautier O, Kaltschmidt JA, Greenleaf WJ,
Gitler AD.Single-cell transcriptomicanalysis of the adult mouse spi-
nal cord reveals molecular diversity of autonomic and skeletal
motor neurons. Nat Neurosci 24: 572583, 2021. doi:10.1038/
s41593-020-00795-0.
327. Siddique U, Rahman S, Frazer AK, Pearce AJ, Howatson G, Kidgell
DJ.Determining the sites of neural adaptations to resistance train-
ing: a systematic review and meta-analysis. Sports Med 50: 1107
1128, 2020. doi:10.1007/s40279-020-01258-z.
328. Kidgell DJ, Bonanno DR, Frazer AK, Howatson G, Pearce AJ.
Corticospinal responses following strength training: a systematic
review and meta-analysis. Eur J Neurosci 46: 26482661, 2017.
doi:10.1111/ejn.13710.
329.
Skarabot J, Brownstein CG, Casolo A, Del Vecchio A, Ansdell P.
The knowns and unknowns of neural adaptations to resistance
training. Eur J Appl Physiol 121: 675685, 2021. doi:10.1007/
s00421-020-04567-3.
330. Del Vecchio A, Negro F, Holobar A, Casolo A, Folland JP, Felici F,
Farina D.You are as fast as your motor neurons: speed of recruit-
ment and maximal discharge of motor neurons determine the maxi-
mal rate of force development in humans. J Physiol 597: 2445
2456, 2019. doi:10.1113/JP277396.
331. Del Vecchio A, Casolo A, Negro F, Scorcelletti M, Bazzucchi I,
Enoka R, Felici F, Farina D.The increase in muscle force after 4
weeks of strength training is mediated by adaptations in motor unit
recruitment and rate coding. J Physiol 597: 18731887, 2019.
doi:10.1113/JP277250.
332. Cormie P, McGuigan MR, Newton RU.Developing maximal neu-
romuscular power: Part 1biological basis of maximal power pro-
duction. Sports Med 41: 1738, 2011. doi:10.2165/11537690-
000000000-00000.
333. Folland JP, Williams AG.The adaptations to strength training:
morphological and neurological contributions to increased
strength. Sports Med 37: 145168, 2007. doi:10.2165/00007256-
200737020-00004.
334. Amiridis IG, Martin A, Morlon B, Martin L, Cometti G, Pousson M, van
Hoecke J.Co-activation and tension-regulating phenomena during
isokinetic knee extension in sedentary and highly skilled humans.
Eur J Appl Physiol Occup Physiol 73: 149156, 1996. doi:10.1007/
BF00262824.
335. Seynnes OR, de Boer M, Narici MV.Early skeletal muscle hyper-
trophy and architectural changes in response to high-intensity
resistance training. JApplPhysiol(1985)102: 368373, 2007.
doi:10.1152/japplphysiol.00789.2006.
336. McKendry J, Stokes T, McLeod JC, Phillips SM.Resistance exercise,
aging, disuse, and muscle protein metabolism. Compr Physiol 11:
22492278, 2021. doi:10.1002/cphy.c200029.
337. Ziliaskoudis C, Park SY, Lee SH.Running economyacomprehen-
sive review for passive force generation. J Exerc Rehabil 15: 640
646, 2019. doi:10.12965/jer.1938406.203.
338. Barnes KR, Kilding AE.Running economy: measurement, norms,
and determining factors. Sports Med Open 1: 8, 2015. doi:10.1186/
s40798-015-0007-y.
339. Hunter SK.Performance fatigability: mechanisms and task specic-
ity. Cold Spring Harb Perspect Med 8: a029728, 2018. doi:10.1101/
cshperspect.a029728.
340. Fukutani A, Tsuruhara Y, Miyake Y, Takao K, Ueno H, Otsuka M,
Suga T, Terada M, Nagano A, Isaka T.Comparison of the relative
muscle volume of triceps surae among sprinters, runners, and
untrained participants. Physiol Rep 8: e14588, 2020. doi:10.14814/
phy2.14588.
341. Methenitis S, Karandreas N, Spengos K, Zaras N, Stasinaki AN,
Terzis G.Muscle ber conduction velocity, muscle ber composi-
tion, and power performance. Med Sci Sports Exerc 48: 17611771,
2016. doi:10.1249/MSS.0000000000000954.
342. Jorgenson KW, Phillips SM, Hornberger TA.Identifying the struc-
tural adaptations that drive the mechanical load-induced growth
of skeletal muscle: a scoping review. Cells 9: 1658, 2020.
doi:10.3390/cells9071658.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1769
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
343. Ema R, Wakahara T, Yanaka T, Kanehisa H, Kawakami Y.Unique
muscularity in cycliststhigh and trunk: a cross-sectional and longi-
tudinal study. Scand J Med Sci Sports 26: 782793, 2016.
doi:10.1111/sms.12511.
344. Kawakami Y, Abe T, Kanehisa H, Fukunaga T.Human skeletal mus-
cle size and architecture: variability and interdependence. Am J
Hum Biol 18: 845848, 2006. doi:10.1002/ajhb.20561.
345. Abe T, Kumagai K, Brechue WF.Fascicle length of leg muscles is
greater in sprinters than distance runners. Med Sci Sports Exerc
32: 11251129, 2000. doi:10.1097/00005768-200006000-00014.
346. Lee HJ, Lee KW, Takeshi K, Lee YW, Kim HJ.Correlation analysis
between lower limb muscle architectures and cycling power via
ultrasonography. Sci Rep 11: 5362, 2021. doi:10.1038/s41598-021-
84870-x.
347. Charles J, Kissane R, Hoehfurtner T, Bates KT.From bre to func-
tion: are we accurately representing muscle architecture and per-
formance? Biol Rev Camb Philos Soc 97: 16401676, 2022.
doi:10.1111/brv.12856.
348. Lieber RL.Can we just forget about pennation angle? JBiomech
132: 110954, 2022. doi:10.1016/j.jbiomech.2022.110954.
349. Roberts MD, Haun CT, Vann CG, Osburn SC, Young KC.
Sarcoplasmic hypertrophy in skeletal muscle: a scienticunicorn
or resistance training adaptation? Front Physiol 11: 816, 2020.
doi:10.3389/fphys.2020.00816.
350. Aagaard P, Andersen JL, Dyhre-Poulsen P, Leffers AM, Wagner A,
Magnusson SP, Halkjaer-Kristensen J, Simonsen EB.A mechanism for
increased contractile strength of human pennate muscle in response
to strength training: changes in muscle architecture. JPhysiol534:
613623, 2001. doi:10.1111/j.1469-7793.2001.t01-1-00613.x.
351. Nielsen J, Ørtenblad N.Physiological aspects of the subcellular
localization of glycogen in skeletal muscle. Appl Physiol Nutr
Metab 38: 9199, 2013. doi:10.1139/apnm-2012-0184.
352. Dankel SJ, Kang M, Abe T, Loenneke JP.Resistance training
induced changes in strength and specic force at the ber and
whole muscle level: a meta-analysis. Eur J Appl Physiol 119: 265
278, 2019. doi:10.1007/s00421-018-4022-9.
353. Kumagai K, Abe T, Brechue WF, Ryushi T, Takano S, Mizuno M.
Sprint performance is related to muscle fascicle length in male 100-
msprinters.J Appl Physiol (1985) 88: 811816, 2000. doi:10.1152/
jappl.2000.88.3.811.
354. Nasirzade A, Ehsanbakhsh A, Ilbeygi S, Sobhkhiz A, Argavani H,
Aliakbari M.Relationship between sprint performance of front crawl
swimming and muscle fascicle length in young swimmers. JSports
Sci Med 13: 550556, 2014.
355. Willingham TB, Kim Y, Lindberg E, Bleck CK, Glancy B.The unied
myobrillar matrix for force generation in muscle. Nat Commun 11:
3722, 2020. doi:10.1038/s41467-020-17579-6.
356. Reggiani C, Schiafno S.Muscle hypertrophy and muscle strength:
dependent or independent variables? A provocative review. Eur J
Transl Myol 30: 9311, 2020. doi:10.4081/ejtm.2020.9311.
357. Schiafno S, Reggiani C.Fiber types in mammalian skeletal
muscles. Physiol Rev 91: 14471531, 2011. doi:10.1152/physrev.
00031.2010.
358. Prasad V, Millay DP.Skeletal muscle bers count on nuclear num-
bers for growth. Semin Cell Dev Biol 119: 310, 2021. doi:10.1016/j.
semcdb.2021.04.015.
359. Hansson KA, Eftestøl E, Bruusgaard JC, Juvkam I, Cramer AW,
Malthe-Sørenssen A, Millay DP, Gundersen K.Myonuclear content
regulates cell size with similar scaling properties in mice and
humans. Nat Commun 11: 6288, 2020. doi:10.1038/s41467-020-
20057-8.
360. Beiter T, Niess AM, Moser D.Transcriptional memory in skeletal
muscle. Dontforget(to)exercise.J Cell Physiol 235: 54765489,
2020. doi:10.1002/jcp.29535.
361. Psilander N, Eftestøl E, Cumming KT, Juvkam I, Ekblom MM,
Sunding K, Wernbom M, Holmberg HC, Ekblom B, Bruusgaard JC,
Raastad T, Gundersen K.Effects of training, detraining, and retrain-
ing on strength, hypertrophy, and myonuclear number in human
skeletal muscle. J Appl Physiol (1985) 126: 16361645, 2019.
doi:10.1152/japplphysiol.00917.2018.
362. Conceição MS, Vechin FC, Lixandrão M, Damas F, Libardi CA,
Tricoli V, Roschel H, Camera D, Ugrinowitsch C.Muscle ber hyper-
trophy and myonuclei addition: a systematic review and meta-anal-
ysis. Med Sci Sports Exerc 50: 13851393, 2018. doi:10.1249/
MSS.0000000000001593.
363. Murach KA, Fry CS, Kirby TJ, Jackson JR, Lee JD, White SH,
Dupont-Versteegden EE, McCarthy JJ, Peterson CA.Starring or
supporting role? Satellite cells and skeletal muscle ber size regula-
tion. Physiology (Bethesda) 33: 2638, 2018. doi:10.1152/physiol.
00019.2017.
364. Kadi F, Eriksson A, Holmner S, Butler-Browne GS, Thornell LE.
Cellular adaptation of the trapezius muscle in strength-trained
athletes. Histochem Cell Biol 111: 189195, 1999. doi:10.1007/
s004180050348.
365. Eriksson A, Kadi F, Malm C, Thornell LE.Skeletal muscle morphol-
ogy in power-lifters with and without anabolic steroids. Histochem
Cell Biol 124: 167175, 2005. doi:10.1007/s00418-005-0029-5.
366. Cramer AA, Prasad V, Eftestøl E, Song T, Hansson KA, Dugdale
HF, Sadayappan S, Ochala J, Gundersen K, Millay DP.Nuclear
numbers in syncytial muscle bers promote size but limit the de-
velopment of larger myonuclear domains. Nat Commun 11: 6287,
2020. doi:10.1038/s41467-020-20058-7.
367. Tillin NA, Jimenez-Reyes P, Pain MT, Folland JP.Neuromuscular
performance of explosive power athletes versus untrained individu-
als. Med Sci Sports Exerc 42: 781790, 2010. doi:10.1249/MSS.
0b013e3181be9c7e.
368. Fry AC, Webber JM, Weiss LW, Harber MP, Vaczi M, Pattison NA.
Muscle ber characteristics of competitive power lifters. J
Strength Cond Res 17: 402410, 2003. doi:10.1519/1533-4287
(2003)017<0402:mfcocp>2.0.co;2.
369. Zaras N, Stasinaki AN, Spiliopoulou P, Arnaoutis G, Hadjicharalambous
M, Terzis G.Rate of force development, muscle architecture, and per-
formance in elite weightlifters. Int J Sports Physiol Perform 16: 216
223, 2021. doi:10.1123/ijspp.2019-0974.
370. Greising SM, Gransee HM, Mantilla CB, Sieck GC.Systems biology
of skeletal muscle: ber type as an organizing principle. Wiley
Interdiscip Rev Syst Biol Med 4: 457473, 2012. doi:10.1002/
wsbm.1184.
371. Qaisar R, Bhaskaran S, Van Remmen H.Muscle ber type diversi-
cation during exercise and regeneration. Free Radic Biol Med 98:
5667, 2016. doi:10.1016/j.freeradbiomed.2016.03.025.
372. Plotkin DL, Roberts MD, Haun CT, Schoenfeld BJ.Muscle ber type
transitions with exercise training: shifting perspectives. Sports 9:
127, 2021. doi:10.3390/sports9090127.
FURRER ET AL.
1770 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
373. Percario V, Boncompagni S, Protasi F, Pertici I, Pinzauti F, Caremani
M.Mechanical parameters of the molecular motor myosin II deter-
mined in permeabilised bres from slow and fast skeletal muscles
of the rabbit. J Physiol 596: 12431257, 2018. doi:10.1113/JP275404.
374. Hvid LG, Gejl K, Bech RD, Nygaard T, Jensen K, Frandsen U,
Ørtenblad N.Transient impairments in single muscle bre contract-
ile function after prolonged cycling in elite endurance athletes.
Acta Physiol (Oxf) 208: 265273, 2013. doi:10.1111/apha.12095.
375. Hvid LG, Ortenblad N, Aagaard P, Kjaer M, Suetta C.Effects of age-
ing on single muscle bre contractile function following short-term
immobilisation. J Physiol 589: 47454757, 2011. doi:10.1113/jphysiol.
2011.215434.
376. Lamboley CR, Rouffet DM, Dutka TL, McKenna MJ, Lamb GD.
Effects of high-intensity intermittent exercise on the contractile
properties of human type I and type II skeletal muscle bers. JAppl
Physiol (1985) 128: 12071216, 2020. doi:10.1152/japplphysiol.
00014.2020.
377. Mitchell EA, Martin NRW, Bailey SJ, Ferguson RA.Critical power is
positively related to skeletal muscle capillarity and type I muscle
bers in endurance-trained individuals. J Appl Physiol (1985) 125:
737745, 2018. doi:10.1152/japplphysiol.01126.2017.
378. Edwards R, Young A, Wiles M.Needle biopsy of skeletal muscle in
the diagnosis of myopathy and the clinical study of muscle function
and repair. N Engl J Med 302: 261271, 1980. doi:10.1056/
NEJM198001313020504.
379. Ball-Burnett M, Green HJ, Houston ME.Energy metabolism in
human slow and fast twitch bres during prolonged cycle exercise.
J Physiol 437: 257267, 1991. doi:10.1113/jphysiol.1991.sp018594.
380. Nielsen J, Suetta C, Hvid LG, Schrøder HD, Aagaard P, Ortenblad
N.Subcellular localization-dependent decrements in skeletal mus-
cle glycogen and mitochondria content following short-term disuse
in young and old men. Am J Physiol Endocrinol Metab 299:
E1053E1060, 2010. doi:10.1152/ajpendo.00324.2010.
381. De Bock K, Derave W, Ramaekers M, Richter EA, Hespel P.Fiber
type-specic muscle glycogen sparing due to carbohydrate intake
before and during exercise. J Appl Physiol (1985) 102: 183188,
2007. doi:10.1152/japplphysiol.00799.2006.
382. Greenhaff PL, Nevill ME, Soderlund K, Bodin K, Boobis LH, Williams
C, Hultman E.The metabolic responses of human type I and II mus-
cle bres during maximal treadmill sprinting. J Physiol 478: 149
155, 1994. doi:10.1113/jphysiol.1994.sp020238.
383. Tsintzas OK, Williams C, Boobis L, Greenhaff P.Carbohydrate
ingestion and single muscle ber glycogen metabolism during pro-
longed running in men. J Appl Physiol (1985) 81: 801809, 1996.
doi:10.1152/jappl.1996.81.2.801.
384. Albers PH, Pedersen AJ, Birk JB, Kristensen DE, Vind BF, Baba O,
Nøhr J, Højlund K, Wojtaszewski JF.Human muscle ber type-spe-
cic insulin signaling: impact of obesity and type 2 diabetes.
Diabetes 64: 485497, 2015. doi:10.2337/db14-0590.
385. Bell DG, Jacobs I.Muscle ber-specic glycogen utilization in
strength-trained males and females. Med Sci Sports Exerc 21: 649
654, 1989. doi:10.1249/00005768-198912000-00004.
386. Gejl KD, Ørtenblad N, Andersson E, Plomgaard P, Holmberg HC,
Nielsen J.Local depletion of glycogen with supramaximal exercise
in human skeletal muscle bres. J Physiol 595: 28092821, 2017.
doi:10.1113/JP273109.
387. Johnson MA, Polgar J, Weightman D, Appleton D.Data on the dis-
tribution of bre types in thirty-six human muscles. An autopsy
study. JNeurolSci18: 111129, 1973. doi:10.1016/0022-510x(73)
90023-3.
388. Horwath O, Envall H, R
oja J, Emanuelsson EB, Sanz G, Ekblom B,
Apr
oW,MobergM.Variability in vastus lateralis ber type distribu-
tion, ber size, and myonuclear content along and between the
legs. J Appl Physiol (1985) 131: 158173, 2021. doi:10.1152/
japplphysiol.00053.2021.
389. Fink WJ, Costill DL, Pollock ML.Submaximal and maximal working
capacity of elite distance runners. Part II. Muscle ber composition
and enzyme activities. Ann NY Acad Sci 301: 323327, 1977.
doi:10.1111/j.1749-6632.1977.tb38210.x.
390. Serrano N, Colenso-Semple LM, Lazauskus KK, Siu JW, Bagley JR,
Lockie RG, Costa PB, Galpin AJ.Extraordinary fast-twitch ber
abundance in elite weightlifters. PLoS One 14: e0207975, 2019.
doi:10.1371/journal.pone.0207975.
391. Trappe S, Luden N, Minchev K, Raue U, Jemiolo B, Trappe TA.
Skeletal muscle signature of a champion sprint runner. JAppl
Physiol (1985) 118: 14601466, 2015. doi:10.1152/japplphysiol.00037.
2015.
392. Baker JS, McCormick MC, Robergs RA.Interaction among skeletal
muscle metabolic energy systems during intense exercise. JNutr
Metab 2010: 905612, 2010. doi:10.1155/2010/905612.
393. Hargreaves M, Spriet LL.Skeletal muscle energy metabolism during
exercise. Nat Metab 2: 817828, 2020. doi:10.1038/s42255-020-
0251-4.
394. Ruby BC, Cuddy JS, Hailes WS, Dumke CL, Slivka DR, Shriver TC,
Schoeller DA.Extreme endurance and the metabolic range of sus-
tained activity is uniquely available for every human not just the elite
few. Comp Exerc Physiol 11: 17, 2015. doi:10.3920/CEP140025.
395. Holloszy JO, Rennie MJ, Hickson RC, Conlee RK, Hagberg JM.
Physiological consequences of the biochemical adaptations to en-
durance exercise. Ann NY Acad Sci 301: 440450, 1977. doi:10.1111/
j.1749-6632.1977.tb38220.x.
396. MacRae HS, Dennis SC, Bosch AN, Noakes TD.Effects of training
on lactate production and removal during progressive exercise in
humans. J Appl Physiol (1985) 72: 16491656, 1992. doi:10.1152/
jappl.1992.72.5.1649.
397. Burke LM.Ketogenic low-CHO, high-fat diet: the future of elite endur-
ance sport? JPhysiol599: 819843, 2021. doi:10.1113/JP278928.
398. Romijn JA, Coyle EF, Sidossis LS, Zhang XJ, Wolfe RR.Relationship
between fatty acid delivery and fatty acid oxidation during strenu-
ous exercise. J Appl Physiol (1985) 79: 19391945, 1995.
doi:10.1152/jappl.1995.79.6.1939.
399. Drake JC, Wilson RJ, Yan Z.Molecular mechanisms for mitochon-
drial adaptation to exercise training in skeletal muscle. FASEB J 30:
1322, 2016. doi:10.1096/fj.15-276337.
400. Bo H, Zhang Y, Ji LL.Redening the role of mitochondria in exer-
cise: a dynamic remodeling. Ann NY Acad Sci 1201: 121128, 2010.
doi:10.1111/j.1749-6632.2010.05618.x.
401. Liesa M, Palacín M, Zorzano A.Mitochondrial dynamics in mamma-
lian health and disease. Physiol Rev 89: 799845, 2009.
doi:10.1152/physrev.00030.2008.
402. Romanello V, Sandri M.The connection between the dynamic
remodeling of the mitochondrial network and the regulation of mus-
cle mass. Cell Mol Life Sci 78: 13051328, 2021. doi:10.1007/
s00018-020-03662-0.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1771
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
403. Kleele T, Rey T, Winter J, Zaganelli S, Mahecic D, Perreten Lambert
H, Ruberto FP, Nemir M, Wai T, Pedrazzini T, Manley S.Distinct s-
sion signatures predict mitochondrial degradation or biogenesis.
Nature 593: 435439, 2021. doi:10.1038/s41586-021-03510-6.
404. Djalalvandi A, Scorrano L.Mitochondrial dynamics: roles in exercise
physiology and muscle mass regulation. Curr Opin Physiol 27:
100550, 2022. doi:10.1016/j.cophys.2022.100550.
405. Perry CG, Lally J, Holloway GP, Heigenhauser GJ, Bonen A, Spriet LL.
Repeated transient mRNA bursts precede increases in transcriptional
and mitochondrial proteins during training in human skeletal muscle.
JPhysiol588: 47954810, 2010. doi:10.1113/jphysiol.2010.199448.
406. Jamart C, Francaux M, Millet GY, Deldicque L, Fr
ere D, F
easson L.
Modulation of autophagy and ubiquitin-proteasome pathways dur-
ing ultra-endurance running. J Appl Physiol (1985) 112: 15291537,
2012. doi:10.1152/japplphysiol.00952.2011.
407. Mishra P, Varuzhanyan G, Pham AH, Chan DC.mitochondrial dy-
namics is a distinguishing feature of skeletal muscle ber types and
regulates organellar compartmentalization. Cell Metab 22: 1033
1044, 2015. doi:10.1016/j.cmet.2015.09.027.
408. Glancy B, Balaban RS.Energy metabolism design of the striated
muscle cell. Physiol Rev 101: 15611607, 2021. doi:10.1152/physrev.
00040.2020.
409. Bleck CK, Kim Y, Willingham TB, Glancy B.Subcellular connectomic
analyses of energy networks in striated muscle. Nat Commun 9:
5111, 2018. doi:10.1038/s41467-018-07676-y.
410. Dahl R, Larsen S, Dohlmann TL, Qvortrup K, Helge JW, Dela F, Prats
C.Three-dimensional reconstruction of the human skeletal muscle
mitochondrial network as a tool to assess mitochondrial content
and structural organization. Acta Physiol (Oxf) 213: 145155, 2015.
doi:10.1111/apha.12289.
411. Vincent AE, White K, Davey T, Philips J, Ogden RT, Lawless C,
Warren C, Hall MG, Ng YS, Falkous G, Holden T, Deehan D, Taylor
RW, Turnbull DM, Picard M.Quantitative 3D mapping of the human
skeletal muscle mitochondrial network. Cell Rep 26: 9961009.e4,
2019. doi:10.1016/j.celrep.2019.01.010.
412. Glancy B, Hartnell LM, Malide D, Yu ZX, Combs CA, Connelly PS,
Subramaniam S, Balaban RS.Mitochondrial reticulum for cellular
energy distribution in muscle. Nature 523: 617620, 2015.
doi:10.1038/nature14614.
413. Lundby C, Jacobs RA.Adaptations of skeletal muscle mitochondria
to exercise training. Exp Physiol 101: 1722, 2016. doi:10.1113/
EP085319.
414. Howald H, Hoppeler H, Claassen H, Mathieu O, Straub R.
Inuences of endurance training on the ultrastructural composition
of the different muscle ber types in humans. Pugers Arch 403:
369376, 1985. doi:10.1007/BF00589248.
415. Jacobs RA, Lundby C.Mitochondria express enhanced quality as
well as quantity in association with aerobic tness across recrea-
tionally active individuals up to elite athletes. J Appl Physiol (1985)
114: 344350, 2013. doi:10.1152/japplphysiol.01081.2012.
416. Granata C, Jamnick NA, Bishop DJ.Training-induced changes in mi-
tochondrial content and respiratory function in human skeletal mus-
cle. Sports Med 48: 18091828, 2018. doi:10.1007/s40279-018-
0936-y.
417. Daemen S, Gemmink A, Brouwers B, Meex RC, Huntjens PR,
Schaart G, Moonen-Kornips E, J
orgensen J, Hoeks J, Schrauwen P,
Hesselink MK.Distinct lipid droplet characteristics and distribution
unmask the apparent contradiction of the athletesparadox.Mol
Metab 17: 7181, 2018. doi:10.1016/j.molmet.2018.08.004.
418. Hoppeler H, Fluck M.Plasticity of skeletal muscle mitochondria:
structure and function. Med Sci Sports Exerc 35: 95104, 2003.
doi:10.1249/01.MSS.0000043292.99104.12.
419. van Loon LJ, Koopman R, Manders R, van der Weegen W, van
Kranenburg GP, Keizer HA.Intramyocellular lipid content in type 2
diabetes patients compared with overweight sedentary men and
highly trained endurance athletes. Am J Physiol Endocrinol Metab
287: E558E565, 2004. doi:10.1152/ajpendo.00464.2003.
420. van Loon LJ, Koopman R, Stegen JH, Wagenmakers AJ, Keizer HA,
Saris WH.Intramyocellular lipids form an important substrate source
during moderate intensity exercise in endurance-trained males in a
fasted state. J Physiol 553: 611625, 2003. doi:10.1113/jphysiol.
2003.052431.
421. Nielsen J, Mogensen M, Vind BF, Sahlin K, Højlund K, Schrøder HD,
Ortenblad N.Increased subsarcolemmal lipids in type 2 diabetes:
effect of training on localization of lipids, mitochondria, and glyco-
gen in sedentary human skeletal muscle. Am J Physiol Endocrinol
Metab 298: E706E713, 2010. doi:10.1152/ajpendo.00692.2009.
422. Koh HE, Nielsen J, Saltin B, Holmberg HC, Ørtenblad N.
Pronounced limb and bre type differences in subcellular lipid
droplet content and distribution in elite skiers before and after ex-
haustive exercise. J Physiol 595: 57815795, 2017. doi:10.1113/
JP274462.
423. Li X, Li Z, Zhao M, Nie Y, Liu P, Zhu Y, Zhang X.Skeletal muscle lipid
droplets and the athletesparadox.Cells 8: 249, 2019. doi:10.3390/
cells8030249.
424. Nielsen J, Holmberg HC, Schrøder HD, Saltin B, Ortenblad N.
Human skeletal muscle glycogen utilization in exhaustive exercise:
role of subcellular localization and bre type. J Physiol 589: 2871
2885, 2011. doi:10.1113/jphysiol.2010.204487.
425. Nielsen J, Schrøder HD, Rix CG, Ortenblad N.Distinct effects of
subcellular glycogen localization on tetanic relaxation time and en-
durance in mechanically skinned rat skeletal muscle bres. J
Physiol 587: 36793690, 2009. doi:10.1113/jphysiol.2009.174862.
426. Hawley JA, Schabort EJ, Noakes TD, Dennis SC.Carbohydrate-
loading and exercise performance. An update. Sports Med 24: 73
81, 1997. doi:10.2165/00007256-199724020-00001.
427. Holloszy JO.Regulation by exercise of skeletal muscle content of
mitochondria and GLUT4. J Physiol Pharmacol 59, Suppl 7: 518,
2008.
428. Burke LM, van Loon LJ, Hawley JA.Postexercise muscle glycogen
resynthesis in humans. J Appl Physiol (1985) 122: 10551067, 2017.
doi:10.1152/japplphysiol.00860.2016.
429. Staron RS, Leonardi MJ, Karapondo DL, Malicky ES, Falkel JE,
Hagerman FC, Hikida RS.Strength and skeletal muscle adaptations
in heavy-resistance-trained women after detraining and retraining.
J Appl Physiol (1985) 70: 631640, 1991. doi:10.1152/jappl.1991.
70.2.631.
430. Seaborne RA, Strauss J, Cocks M, Shepherd S, OBrien TD, van
Someren KA, Bell PG, Murgatroyd C, Morton JP, Stewart CE,
Sharples AP.Human skeletal muscle possesses an epigenetic
memory of hypertrophy. Sci Rep 8: 1898, 2018. doi:10.1038/s41598-
018-20287-3.
431. Wen Y, Dungan CM, Mobley CB, Valentino T, von Walden F,
Murach KA.Nucleus type-specic DNA methylomics reveals
FURRER ET AL.
1772 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
epigenetic memoryof prior adaptation in skeletal muscle.
Function 2: zqab038, 2021. doi:10.1093/function/zqab038.
432. Egner IM, Bruusgaard JC, Eftestøl E, Gundersen K.A cellular mem-
ory mechanism aids overload hypertrophy in muscle long after an
episodic exposure to anabolic steroids. J Physiol 591: 62216230,
2013. doi:10.1113/jphysiol.2013.264457.
433. Snijders T, Aussieker T, Holwerda A, Parise G, van Loon LJ, Verdijk
LB.The concept of skeletal muscle memory: evidence from animal
and human studies. Acta Physiol (Oxf) 229: e13465, 2020.
doi:10.1111/apha.13465.
434. Murach KA, Dungan CM, Dupont-Versteegden EE, McCarthy JJ,
Peterson CA.Muscle memorynot mediated by myonuclear num-
ber? Secondary analysis of human detraining data. J Appl Physiol
(1985) 127: 18141816, 2019. doi:10.1152/japplphysiol.00506.2019.
435. Sharples AP, Stewart CE, Seaborne RA.Does skeletal muscle have
an epi-memory? The role of epigenetics in nutritional program-
ming, metabolic disease, aging and exercise. Aging Cell 15: 603
616, 2016. doi:10.1111/acel.12486.
436. Cedar H, Bergman Y.Linking DNA methylation and histone modi-
cation: patterns and paradigms. Nat Rev Genet 10: 295304,
2009. doi:10.1038/nrg2540.
437. Zhang T, Cooper S, Brockdorff N.The interplay of histone modica-
tionswriters that read. EMBO Rep 16: 14671481, 2015. doi:10.15252/
embr.201540945.
438. Sharples AP, Seaborne RA.Exercise and DNA methylation in skele-
tal muscle. In: Sports, Exercise, and Nutritional Genomics: Current Status
and Future Directions,editedbyBarhD,AhmetovII.London:
Academic Press, 2019, p. 211229.
439. Sailani MR, Halling JF, Møller HD, Lee H, Plomgaard P, Pilegaard H,
Snyder MP, Regenberg B.Lifelong physical activity is associated
with promoter hypomethylation of genes involved in metabolism,
myogenesis, contractile properties and oxidative stress resistance
in aged human skeletal muscle. Sci Rep 9: 3272, 2019. doi:10.1038/
s41598-018-37895-8.
440. Alibegovic AC, Sonne MP, Højbjerre L, Bork-Jensen J, Jacobsen S,
Nilsson E, Faerch K, Hiscock N, Mortensen B, Friedrichsen M,
Stallknecht B, Dela F, Vaag A.Insulin resistance induced by physi-
cal inactivity is associated with multiple transcriptional changes in
skeletal muscle in young men. Am J Physiol Endocrinol Metab
299: E752E763, 2010. doi:10.1152/ajpendo.00590.2009.
441. Barr
es R, Yan J, Egan B, Treebak JT, Rasmussen M, Fritz T, Caidahl
K, Krook A, OGorman DJ, Zierath JR.Acute exercise remodels pro-
moter methylation in human skeletal muscle. Cell Metab 15: 405
411, 2012. doi:10.1016/j.cmet.2012.01.001.
442. Nitert MD, Dayeh T, Volkov P, Elgzyri T, Hall E, Nilsson E, Yang BT,
Lang S, Parikh H, Wessman Y, Weishaupt H, Attema J, Abels M,
Wierup N, Almgren P, Jansson PA, R
onn T, Hansson O, Eriksson
KF, Groop L, Ling C.Impact of an exercise intervention on DNA
methylation in skeletal muscle from rst-degree relatives of patients
with type 2 diabetes. Diabetes 61: 33223332, 2012. doi:10.2337/
db11-1653.
443. Lindholm ME, Marabita F, Gomez-Cabrero D, Rundqvist H, Ekstr
om
TJ, Tegn
er J, Sundberg CJ.An integrative analysis reveals coordi-
nated reprogramming of the epigenome and the transcriptome in
human skeletal muscle after training. Epigenetics 9: 15571569,
2014. doi:10.4161/15592294.2014.982445.
444. Robinson MM, Dasari S, Konopka AR, Johnson ML, Manjunatha S,
Esponda RR, Carter RE, Lanza IR, Nair KS.Enhanced protein transla-
tion underlies improved metabolic and physical adaptations to
different exercise training modes in young and old humans. Cell
Metab 25: 581592, 2017. doi:10.1016/j.cmet.2017.02.009.
445. Telles GD, Libardi CA, Conceição MS, Vechin FC, Lixandrão ME,
Mangone FR, Pavanelli AC, Nagai MA, Camera DM, Hawley JA,
Ugrinowitsch C.Interrelated but not time-aligned response in myo-
genic regulatory factors demethylation and mRNA expression after
divergent exercise bouts. Med Sci Sports Exerc 55: 199208,
2023. doi:10.1249/MSS.0000000000003049.
446. Kuang J, McGinley C, Lee MJ, Saner NJ, Garnham A, Bishop DJ.
Interpretation of exercise-induced changes in human skeletal mus-
cle mRNA expression depends on the timing of the post-exercise
biopsies. PeerJ 10: e12856, 2022. doi:10.7717/peerj.12856.
447. Turner DC, Seaborne RA, Sharples AP.Comparative transcriptome
and methylome analysis in human skeletal muscle anabolism, hy-
pertrophy and epigenetic memory. Sci Rep 9: 4251, 2019.
doi:10.1038/s41598-019-40787-0.
448. Jacques M, Hiam D, Craig J, Barr
es R, Eynon N, Voisin S.
Epigenetic changes in healthy human skeletal muscle following
exercisea systematic review. Epigenetics 14: 633648, 2019.
doi:10.1080/15592294.2019.1614416.
449. Basang Z, Zhang S, Yang L, Quzong D, Li Y, Ma Y, Hao M, Pu WL,
Liu X, Xie H, Liang M, Wang J, Danzeng Q.Correlation of DNA
methylation patterns to the phenotypic features of Tibetan elite
alpinists in extreme hypoxia. J Genet Genomics 48: 928935,
2021. doi:10.1016/j.jgg.2021.05.015.
450. Sp
olnicka M, Po
spiech E, Adamczyk JG, Freire-Aradas A, Pepło
nska
B, Zbieć-Piekarska R, Makowska _
Z, Pięta A, Lareu MV, Phillips C,
Płoski R, _
Zekanowski C, Branicki W.Modied aging of elite athletes
revealed by analysis of epigenetic age markers. Aging (Albany NY)
10: 241252, 2018. doi:10.18632/aging.101385.
451. Terruzzi I, Senesi P, Montesano A, La Torre A, Alberti G, Benedini S,
Caumo A, Fermo I, Luzi L.Genetic polymorphisms of the enzymes
involved in DNA methylation and synthesis in elite athletes. Physiol
Genomics 43: 965973, 2011. doi:10.1152/physiolgenomics.00040.
2010.
452. Bishop DJ, Hawley JA.Reassessing the relationship between
mRNA levels and protein abundance in exercised skeletal muscles.
Nat Rev Mol Cell Biol 23: 773774, 2022. doi:10.1038/s41580-022-
00541-3.
453. Pillon NJ, Gabriel BM, Dollet L, Smith JA, Sardon Puig L, Botella J,
Bishop DJ, Krook A, Zierath JR.Transcriptomic proling of skeletal
muscle adaptations to exercise and inactivity. Nat Commun 11: 470,
2020. doi:10.1038/s41467-019-13869-w.
454. Amar D, Gay NR, Jean Beltran PM, Adkins JN, Almagro Armenteros
JJ, Ashley E, et al.Temporal dynamics of the multi-omic response
to endurance exercise training across tissues (Preprint). bioRxiv
2022.09.2021.508770, 2022. doi:10.1101/2022.09.21.508770.
455. Gabriel BM, Zierath JR.The limits of exercise physiology: from per-
formance to health. Cell Metab 25: 10001011, 2017. doi:10.1016/j.
cmet.2017.04.018.
456. Akay T, Murray AJ.Relative contribution of proprioceptive and ves-
tibular sensory systems to locomotion: opportunities for discovery
in the age of molecular science. Int J Mol Sci 22: 1467, 2021.
doi:10.3390/ijms22031467.
457. Taylor DF, Bishop DJ.Transcription factor movement and exercise-
induced mitochondrial biogenesis in human skeletal muscle: cur-
rent knowledge and future perspectives. Int J Mol Sci 23: 1517,
2022. doi:10.3390/ijms23031517.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1773
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
458. Roatta S, Farina D.Sympathetic actions on the skeletal mus-
cle. Exerc Sport Sci Rev 38: 3135, 2010. doi:10.1097/JES.
0b013e3181c5cde7.
459. Delbono O, Rodrigues AC, Bonilla HJ, Messi ML.The emerging role
of the sympathetic nervous system in skeletal muscle motor inner-
vation and sarcopenia. Ageing Res Rev 67: 101305, 2021.
doi:10.1016/j.arr.2021.101305.
460. Teixeira AL, Vianna LC.The exercise pressor reex: an update. Clin
Auton Res 32: 271290, 2022. doi:10.1007/s10286-022-00872-3.
461. Joassard OR, Durieux AC, Freyssenet DG.beta2-Adrenergic ago-
nists and the treatment of skeletal muscle wasting disorders. Int J
Biochem Cell Biol 45: 23092321, 2013. doi:10.1016/j.biocel.2013.
06.025.
462. Sandri M, Lin J, Handschin C, Yang W, Arany ZP, Lecker SH,
Goldberg AL, Spiegelman BM.PGC-1alpha protects skeletal muscle
from atrophy by suppressing FoxO3 action and atrophy-specic
gene transcription. Proc Natl Acad Sci USA 103: 1626016265,
2006. doi:10.1073/pnas.0607795103.
463. Schwartz AB.Movement: how the brain communicates with the
world. Cell 164: 11221135, 2016. doi:10.1016/j.cell.2016.02.038.
464. Scott SH.Optimal feedback control and the neural basis of voli-
tional motor control. Nat Rev Neurosci 5: 532546, 2004.
doi:10.1038/nrn1427.
465. Madden CJ, Morrison SF.Central nervous system circuits that con-
trol body temperature. Neurosci Lett 696: 225232, 2019.
doi:10.1016/j.neulet.2018.11.027.
466. Stifani N.Motor neurons and the generation of spinal motor neuron
diversity. Front Cell Neurosci 8: 293, 2014. doi:10.3389/fncel.
2014.00293.
467. Li L, Xiong WC, Mei L.Neuromuscular junction formation, aging,
and disorders. Annu Rev Physiol 80: 159188, 2018. doi:10.1146/
annurev-physiol-022516-034255.
468. Tintignac LA, Brenner HR, Ruegg MA.Mechanisms regulating neu-
romuscular junction development and function and causes of mus-
cle wasting. Physiol Rev 95: 809852, 2015. doi:10.1152/physrev.
00033.2014.
469. Shishmarev D.Excitation-contraction coupling in skeletal muscle:
recent progress and unanswered questions. Biophys Rev 12: 143
153, 2020. doi:10.1007/s12551-020-00610-x.
470. Tu MK, Levin JB, Hamilton AM, Borodinsky LN.Calcium signaling in
skeletal muscle development, maintenance and regeneration. Cell
Calcium 59: 9197, 2016. doi:10.1016/j.ceca.2016.02.005.
471. Gehlert S, Bloch W, Suhr F.Ca
21
-dependent regulations and sig-
naling in skeletal muscle: from electro-mechanical coupling to
adaptation. Int J Mol Sci 16: 10661095, 2015. doi:10.3390/
ijms16011066.
472. Pette D, Vrbová G.The contribution of neuromuscular stimulation in
elucidating muscle plasticity revisited. Eur J Transl Myol 27: 6368,
2017. doi:10.4081/ejtm.2017.6368.
473. Salmons S.The adaptive response of skeletal muscle: what is the
evidence? Muscle Nerve 57: 531541, 2018. doi:10.1002/mus.
25949.
474. Kupr B, Handschin C.Complex coordination of cell plasticity by a
PGC-1alpha-controlled transcriptional network in skeletal muscle.
Front Physiol 6: 325, 2015. doi:10.3389/fphys.2015.00325.
475. Handschin C, Rhee J, Lin J, Tarr PT, Spiegelman BM.An autoregula-
tory loop controls peroxisome proliferator-activated receptor
gamma coactivator 1alpha expression in muscle. Proc Natl Acad
Sci USA 100: 71117116, 2003. doi:10.1073/pnas.1232352100.
476. Czubryt MP, McAnally J, Fishman GI, Olson EN.Regulation of per-
oxisome proliferator-activated receptor gamma coactivator 1 alpha
(PGC-1 alpha) and mitochondrial function by MEF2 and HDAC5.
Proc Natl Acad Sci USA 100: 17111716, 2003. doi:10.1073/pnas.
0337639100.
477. Lin J, Wu H, Tarr PT, Zhang CY, Wu Z, Boss O, Michael LF,
Puigserver P, Isotani E, Olson EN, Lowell BB, Bassel-Duby R,
Spiegelman BM.Transcriptional co-activator PGC-1 alpha drives the
formation of slow-twitch muscle bres. Nature 418: 797801, 2002.
doi:10.1038/nature00904.
478. Arany Z, Foo SY, Ma Y, Ruas JL, Bommi-Reddy A, Girnun G, Cooper
M, Laznik D, Chinsomboon J, Rangwala SM, Baek KH, Rosenzweig
A, Spiegelman BM.HIF-independent regulation of VEGF and angio-
genesis by the transcriptional coactivator PGC-1alpha. Nature 451:
10081012, 2008. doi:10.1038/nature06613.
479. Handschin C, Choi CS, Chin S, Kim S, Kawamori D, Kurpad AJ,
Neubauer N, Hu J, Mootha VK, Kim YB, Kulkarni RN, Shulman GI,
Spiegelman BM.Abnormal glucose homeostasis in skeletal mus-
cle-specic PGC-1alpha knockout mice reveals skeletal muscle-
pancreatic beta cell crosstalk. JClinInvest117: 34633474, 2007.
doi:10.1172/JCI31785.
480. Handschin C, Chin S, Li P, Liu F, Maratos-Flier E, Lebrasseur NK,
Yan Z, Spiegelman BM.Skeletal muscle ber-type switching, exer-
cise intolerance, and myopathy in PGC-1alpha muscle-specic
knock-out animals. JBiolChem282: 3001430021, 2007.
doi:10.1074/jbc.M704817200.
481. Martínez-Redondo V, Pettersson AT, Ruas JL.The hitchhikersguide
to PGC-1alpha isoform structure and biological functions. Diabetologia
58: 19691977, 2015. doi:10.1007/s00125-015-3671-z.
482. Jannig PR, Dumesic PA, Spiegelman BM, Ruas JL.SnapShot:
Regulation and biology of PGC-1alpha. Cell 185: 14441444.e1,
2022. doi:10.1016/j.cell.2022.03.027.
483. P
erez-Schindler J, Kohl B, Schneider-Heieck K, Leuchtmann AB,
Henríquez-Olguín C, Adak V, Maier G, Delezie J, Sakoparnig T,
Vargas-Fernández E, Karrer-Cardel B, Ritz D, Schmidt A, Hondele
M, Jensen TE, Hiller S, Handschin C.RNA-bound PGC-1alpha con-
trols gene expression in liquid-like nuclear condensates. Proc Natl
Acad Sci USA 118: e2105951118, 2021. doi:10.1073/pnas.2105951118.
484. Handschin C, Spiegelman BM.Peroxisome proliferator-activated re-
ceptor gamma coactivator 1 coactivators, energy homeostasis, and
metabolism. Endocr Rev 27: 728735, 2006. doi:10.1210/er.2006-
0037.
485. Kupr B, Schnyder S, Handschin C.Role of nuclear receptors in exer-
cise-induced muscle adaptations. Cold Spring Harb Perspect Med
7: a029835, 2017. doi:10.1101/cshperspect.a029835.
486. Schnyder S, Kupr B, Handschin C.Coregulator-mediated control of
skeletal muscle plasticityamini-review.Biochimie 136: 4954,
2017. doi:10.1016/j.biochi.2016.12.011.
487. Handschin C, Kobayashi YM, Chin S, Seale P, Campbell KP,
Spiegelman BM.PGC-1alpha regulates the neuromuscular junction
program and ameliorates Duchenne muscular dystrophy. Genes
Dev 21: 770783, 2007. doi:10.1101/gad.1525107.
488. Summermatter S, Baum O, Santos G, Hoppeler H, Handschin C.
Peroxisome proliferator-activated receptor gamma coactivator 1alpha
(PGC-1alpha) promotes skeletal muscle lipid refueling in vivo by
FURRER ET AL.
1774 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
activating de novo lipogenesis and the pentose phosphate pathway.
JBiolChem285: 3279332800, 2010. doi:10.1074/jbc.M110.145995.
489. Handschin C.Regulation of skeletal muscle cell plasticity by the per-
oxisome proliferator-activated receptor gamma coactivator 1alpha. J
Recept Signal Transduct Res 30: 376384, 2010. doi:10.3109/
10799891003641074.
490. Amoasii L, Sanchez-Ortiz E, Fujikawa T, Elmquist JK, Bassel-Duby R,
Olson EN.NURR1 activation in skeletal muscle controls systemic
energy homeostasis. Proc Natl Acad Sci USA 116: 1129911308,
2019. doi:10.1073/pnas.1902490116.
491. Goode JM, Pearen MA, Tuong ZK, Wang SC, Oh TG, Shao EX,
Muscat GE.The nuclear receptor, Nor-1, induces the physiological
responses associated with exercise. Mol Endocrinol 30: 660676,
2016. doi:10.1210/me.2015-1300.
492. Mills R, Taylor-Weiner H, Correia JC, Agudelo LZ, Allodi I, Kolonelou
C, Martinez-Redondo V, Ferreira DMS, Nichterwitz S, Comley LH,
Lundin V, Hedlund E, Ruas JL, Teixeira AI.Neurturin is a PGC-
1alpha1-controlled myokine that promotes motor neuron recruit-
ment and neuromuscular junction formation. Mol Metab 7: 1222,
2018. doi:10.1016/j.molmet.2017.11.001.
493. Correia JC, Kelahmetoglu Y, Jannig PR, Schweingruber C,
Shvaikovskaya D, Zhengye L, Cervenka I, Khan N, Stec M, Oliveira
M, Nijssen J, Martínez-Redondo V, Ducommun S, Azzolini M,
Lanner JT, Kleiner S, Hedlund E, Ruas JL.Muscle-secreted neu-
rturin couples myober oxidative metabolism and slow motor neu-
ron identity. Cell Metab 33: 22152230.e8, 2021. doi:10.1016/j.
cmet.2021.09.003.
494. Pedersen BK.Physical activity and muscle-brain crosstalk. Nat Rev
Endocrinol 15: 383392, 2019. doi:10.1038/s41574-019-0174-x.
495. Delezie J, Weihrauch M, Maier G, Tejero R, Ham DJ, Gill JF, Karrer-
Cardel B, R
uegg MA, Tabares L, Handschin C.BDNF is a media-
tor of glycolytic ber-type specication in mouse skeletal muscle.
Proc Natl Acad Sci USA 116: 1611116120, 2019. doi:10.1073/pnas.
1900544116.
496. Steinacker JM, Brkic M, Simsch C, Nething K, Kresz A, Prokopchuk
O, Liu Y.Thyroid hormones, cytokines, physical training and meta-
bolic control. Horm Metab Res 37: 538544, 2005. doi:10.1055/s-
2005-870419.
497. Kraemer WJ, Ratamess NA, Nindl BC.Recovery responses of tes-
tosterone, growth hormone, and IGF-1 after resistance exercise. J
Appl Physiol (1985) 122: 549558, 2017. doi:10.1152/japplphysiol.
00599.2016.
498. Laurens C, Bergouignan A, Moro C.Exercise-released myokines in
the control of energy metabolism. Front Physiol 11: 91, 2020.
doi:10.3389/fphys.2020.00091.
499. Thyfault JP, Bergouignan A.Exercise and metabolic health: beyond
skeletal muscle. Diabetologia 63: 14641474, 2020. doi:10.1007/
s00125-020-05177-6.
500. Chow LS, Gerszten RE, Taylor JM, Pedersen BK, van Praag H,
Trappe S, Febbraio MA, Galis ZS, Gao Y, Haus JM, Lanza IR, Lavie
CJ, Lee CH, Lucia A, Moro C, Pandey A, Robbins JM, Stanford KI,
Thackray AE, Villeda S, Watt MJ, Xia A, Zierath JR, Goodpaster BH,
Snyder MP.Exerkines in health, resilience and disease. Nat Rev
Endocrinol 18: 273289, 2022. doi:10.1038/s41574-022-00641-2.
501. Schnyder S, Handschin C.Skeletal muscle as an endocrine organ:
PGC-1alpha, myokines and exercise. Bone 80: 115125, 2015.
doi:10.1016/j.bone.2015.02.008.
502. Furrer R, Eisele PS, Schmidt A, Beer M, Handschin C.Paracrine
cross-talk between skeletal muscle and macrophages in exercise
by PGC-1alpha-controlled BNP. Sci Rep 7: 40789, 2017. doi:10.1038/
srep40789.
503. Rowe GC, Raghuram S, Jang C, Nagy JA, Patten IS, Goyal A, Chan
MC, Liu LX, Jiang A, Spokes KC, Beeler D, Dvorak H, Aird WC,
Arany Z.PGC-1alpha induces SPP1 to activate macrophages and
orchestrate functional angiogenesis in skeletal muscle. Circ Res
115: 504517, 2014. doi:10.1161/CIRCRESAHA.115.303829.
504. Leuchtmann AB, Adak V, Dilbaz S, Handschin C.The role of the
skeletal muscle secretome in mediating endurance and resistance
training adaptations. Front Physiol 12: 709807, 2021. doi:10.3389/
fphys.2021.709807.
505. Zunner BE, Wachsmuth NB, Eckstein ML, Scherl L, Schierbauer JR,
Haupt S, Stumpf C, Reusch L, Moser O.Myokines and resistance
training: a narrative review. Int J Mol Sci 23: 3501, 2022.
doi:10.3390/ijms23073501.
506. Knudsen NH, Stanya KJ, Hyde AL, Chalom MM, Alexander RK, Liou
YH, Starost KA, Gangl MR, Jacobi D, Liu S, Sopariwala DH,
Fonseca-Pereira D, Li J, Hu FB, Garrett WS, Narkar VA, Ortlund EA,
Kim JH, Paton CM, Cooper JA, Lee CH.Interleukin-13 drives meta-
bolic conditioning of muscle to endurance exercise. Science 368:
eaat3987, 2020. doi:10.1126/science.aat3987.
507. Varga T, Mounier R, Patsalos A, Gogolák P, Peloquin M, Horvath A,
Pap A, Daniel B, Nagy G, Pintye E, P
oliska S, Cuvellier S, Larbi SB,
Sansbury BE, Spite M, Brown CW, Chazaud B, Nagy L.Macrophage
PPARgamma, a lipid activated transcription factor controls the
growth factor GDF3 and skeletal muscle regeneration. Immunity
45: 10381051, 2016. doi:10.1016/j.immuni.2016.10.016.
508. Zhang J, Muri J, Fitzgerald G, Gorski T, Gianni-Barrera R,
Masschelein E, DHulst G, Gilardoni P, Turiel G, Fan Z, Wang T,
Planque M, Carmeliet P, Pellerin L, Wolfrum C, Fendt SM, BanA,
Stockmann C, Soro-Arnáiz I, Kopf M, De Bock K.Endothelial lactate
controls muscle regeneration from ischemia by inducing M2-like
macrophage polarization. Cell Metab 31: 11361153.e7, 2020.
doi:10.1016/j.cmet.2020.05.004.
509. Gonzalez-Gil AM, Elizondo-Montemayor L.The role of exercise in
the interplay between myokines, hepatokines, osteokines, adipo-
kines, and modulation of inammation for energy substrate redis-
tribution and fat mass loss: a review. Nutrients 12: 1899, 2020.
doi:10.3390/nu12061899.
510. Weigert C, Hoene M, Plomgaard P.Hepatokines-a novel group of
exercise factors. Pugers Arch 471: 383396, 2019. doi:10.1007/
s00424-018-2216-y.
511. Burkholder TJ.Mechanotransduction in skeletal muscle. Front
Biosci 12: 174191, 2007. doi:10.2741/2057.
512. Benavides Damm T, Egli M.Calciums role in mechanotransduction
during muscle development. Cell Physiol Biochem 33: 249272,
2014. doi:10.1159/000356667.
513. Bernareggi A, Bosutti A, Massaria G, Giniatullin R, Malm T,
Sciancalepore M, Lorenzon P.The state of the art of Piezo1 chan-
nels in skeletal muscle regeneration. Int J Mol Sci 23: 6616, 2022.
doi:10.3390/ijms23126616.
514. Cullen PJ.Decoding complex Ca
21
signals through the modulation
of Ras signaling. Curr Opin Cell Biol 18: 157161, 2006. doi:10.1016/
j.ceb.2006.02.012.
515. Rindom E, Vissing K.Mechanosensitive molecular networks involved
in transducing resistance exercise-signals into muscle protein accre-
tion. Front Physiol 7: 547, 2016. doi:10.3389/fphys.2016.00547.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1775
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
516. Wright DC, Han DH, Garcia-Roves PM, Geiger PC, Jones TE,
Holloszy JO.Exercise-induced mitochondrial biogenesis begins
before the increase in muscle PGC-1alpha expression. JBiolChem
282: 194199, 2007. doi:10.1074/jbc.M606116200.
517. Kramer HF, Goodyear LJ.Exercise, MAPK, and NF-kappaB signal-
ing in skeletal muscle. J Appl Physiol (1985) 103: 388395, 2007.
doi:10.1152/japplphysiol.00085.2007.
518. Bahrami S, Drabløs F.Gene regulation in the immediate-early
response process. Adv Biol Regul 62: 3749, 2016. doi:10.1016/j.
jbior.2016.05.001.
519. Goodman CA, Hornberger TA, Robling AG.Bone and skeletal mus-
cle: key players in mechanotransduction and potential overlapping
mechanisms. Bone 80: 2436, 2015. doi:10.1016/j.bone.2015.04.014.
520. Attwaters M, Hughes SM.Cellular and molecular pathways control-
ling muscle size in response to exercise. FEBS J 289: 14281456,
2022. doi:10.1111/febs.15820.
521. González A, Hall MN, Lin SC, Hardie DG.AMPK and TOR: the yin
and yang of cellular nutrient sensing and growth control. Cell
Metab 31: 472492, 2020. doi:10.1016/j.cmet.2020.01.015.
522. Yoon MS.mTOR as a key regulator in maintaining skeletal muscle
mass. Front Physiol 8: 788, 2017. doi:10.3389/fphys.2017.00788.
523. Bentzinger CF, Lin S, Romanino K, Castets P, Guridi M, Summermatter
S, Handschin C, Tintignac LA, Hall MN, Ruegg MA.Differential
response of skeletal muscles to mTORC1 signaling during atrophy and
hypertrophy. Skelet Muscle 3: 6, 2013. doi:10.1186/2044-5040-3-6.
524. Ham AS, Chojnowska K, Tintignac LA, Lin S, Schmidt A, Ham DJ,
Sinnreich M, R
uegg MA.mTORC1 signalling is not essential for the
maintenance of muscle mass and function in adult sedentary mice.
J Cachexia Sarcopenia Muscle 11: 259273, 2020. doi:10.1002/
jcsm.12505.
525. Wackerhage H, Schoenfeld BJ, Hamilton DL, Lehti M, Hulmi JJ.
Stimuli and sensors that initiate skeletal muscle hypertrophy follow-
ing resistance exercise. J Appl Physiol (1985) 126: 3043, 2019.
doi:10.1152/japplphysiol.00685.2018.
526. Fischer M, Rikeit P, Knaus P, Coirault C.YAP-mediated mechanotrans-
duction in skeletal muscle. Front Physiol 7: 41, 2016. doi:10.3389/
fphys.2016.00041.
527. Kr
uger M, K
otter S.Titin, a central mediator for hypertrophic signal-
ing, exercise-induced mechanosignaling and skeletal muscle
remodeling. Front Physiol 7: 76, 2016. doi:10.3389/fphys.2016.
00076.
528. Nishikawa K, Lindstedt SL, Hessel A, Mishra D.N2A titin: signaling
hub and mechanical switch in skeletal muscle. Int J Mol Sci 21:
3974, 2020. doi:10.3390/ijms21113974.
529. Ibata N, Terentjev EM.Why exercise builds muscles: titin mechano-
sensing controls skeletal muscle growth under load. Biophys J 120:
36493663, 2021. doi:10.1016/j.bpj.2021.07.023.
530. Freundt JK, Linke WA.Titin as a force-generating muscle protein
under regulatory control. J Appl Physiol (1985) 126: 14741482,
2019. doi:10.1152/japplphysiol.00865.2018.
531. van Ingen MJ, Kirby TJ.LINCing nuclear mechanobiology with skel-
etal muscle mass and function. Front Cell Dev Biol 9: 690577,
2021. doi:10.3389/fcell.2021.690577.
532. Niethammer P.Components and mechanisms of nuclear mechano-
transduction. Annu Rev Cell Dev Biol 37: 233256, 2021.
doi:10.1146/annurev-cellbio-120319-030049.
533. Jabre S, Hleihel W, Coirault C.Nuclear mechanotransduction in
skeletal muscle. Cells 10: 318, 2021. doi:10.3390/cells10020318.
534. Powers SK, Deminice R, Ozdemir M, Yoshihara T, Bomkamp MP,
Hyatt H.Exercise-induced oxidative stress: friend or foe? JSport
Health Sci 9: 415425, 2020. doi:10.1016/j.jshs.2020.04.001.
535. Sies H, Jones DP.Reactive oxygen species (ROS) as pleiotropic
physiological signalling agents. Nat Rev Mol Cell Biol 21: 363383,
2020. doi:10.1038/s41580-020-0230-3.
536. Ward CW, Prosser BL, Lederer WJ.Mechanical stretch-induced acti-
vation of ROS/RNS signaling in striated muscle. Antioxid Redox
Signal 20: 929936, 2014. doi:10.1089/ars.2013.5517.
537. Di Meo S, Napolitano G, Venditti P.Mediators of physical activity
protection against ROS-linked skeletal muscle damage. Int J Mol
Sci 20: 3024, 2019. doi:10.3390/ijms20123024.
538. Nemes R, Koltai E, Taylor AW, Suzuki K, Gyori F, Radak Z.Reactive
oxygen and nitrogen species regulate key metabolic, anabolic, and
catabolic pathways in skeletal muscle. Antioxidants (Basel) 7: 85,
2018. doi:10.3390/antiox7070085.
539. Suhr F, Gehlert S, Grau M, Bloch W.Skeletal muscle function during
exercise-ne-tuning of diverse subsystems by nitric oxide. Int J Mol
Sci 14: 71097139, 2013. doi:10.3390/ijms14047109.
540. Lennicke C, Cochem
eHM.Redox metabolism: ROS as specicmo-
lecular regulators of cell signaling and function. Mol Cell 81: 3691
3707, 2021. doi:10.1016/j.molcel.2021.08.018.
541. Gureev AP, Shaforostova EA, Popov VN.Regulation of mitochon-
drial biogenesis as a way for active longevity: interaction between
the Nrf2 and PGC-1alpha signaling pathways. Front Genet 10: 435,
2019. doi:10.3389/fgene.2019.00435.
542. Beyfuss K, Hood DA.A systematic review of p53 regulation of oxi-
dative stress in skeletal muscle. Redox Rep 23: 100117, 2018.
doi:10.1080/13510002.2017.1416773.
543. Eisele PS, Handschin C.Functional crosstalk of PGC-1 coactivators
and inammation in skeletal muscle pathophysiology. Semin
Immunopathol 36: 2753, 2014. doi:10.1007/s00281-013-0406-4.
544. Nocella C, Cammisotto V, Pigozzi F, Borrione P, Fossati C, DAmico
A, Cangemi R, Peruzzi M, Gobbi G, Ettorre E, Frati G, Cavarretta E,
Carnevale R; SMiLe Group. Impairment between oxidant and anti-
oxidant systems: short- and long-term implications for athletes
health. Nutrients 11: 1353, 2019. doi:10.3390/nu11061353.
545. Knez WL, Jenkins DG, Coombes JS.The effect of an increased
training volume on oxidative stress. Int J Sports Med 35: 813,
2014. doi:10.1055/s-0033-1333746.
546. Knez WL, Jenkins DG, Coombes JS.Oxidative stress in half and full
Ironman triathletes. Med Sci Sports Exerc 39: 283288, 2007.
doi:10.1249/01.mss.0000246999.09718.0c.
547. Varamenti E, Tod D, Pullinger SA.Redox homeostasis and inam-
mation responses to training in adolescent athletes: a systematic
review and meta-analysis. Sports Med Open 6: 34, 2020.
doi:10.1186/s40798-020-00262-x.
548. Pastor R, Tur JA.Antioxidant supplementation and adaptive
response to training: a systematic review. Curr Pharm Des 25:
18891912, 2019. doi:10.2174/1381612825666190701164923.
549. Jackson MJ, Pollock N, Staunton C, Jones S, McArdle A.Redox
control of signalling responses to contractile activity and ageing in
skeletal muscle. Cells 11: 1698, 2022. doi:10.3390/cells11101698.
FURRER ET AL.
1776 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
550. Chini CC, Zeidler JD, Kashyap S, Warner G, Chini EN.Evolving con-
cepts in NAD
1
metabolism. Cell Metab 33: 10761087, 2021.
doi:10.1016/j.cmet.2021.04.003.
551. Glancy B, Kane DA, Kavazis AN, Goodwin ML, Willis WT, Gladden
LB.Mitochondrial lactate metabolism: history and implications for
exercise and disease. J Physiol 599: 863888, 2021. doi:10.1113/
JP278930.
552. Trefts E, Shaw RJ.AMPK: restoring metabolic homeostasis over
space and time. Mol Cell 81: 36773690, 2021. doi:10.1016/j.
molcel.2021.08.015.
553. Spaulding HR, Yan Z.AMPK and the adaptation to exercise. Annu
Rev Physiol 84: 209227, 2022. doi:10.1146/annurev-physiol-
060721-095517.
554. Kjøbsted R, Hingst JR, Fentz J, Foretz M, Sanz MN, Pehmøller C,
Shum M, Marette A, Mounier R, Treebak JT, Wojtaszewski JFP,
Viollet B, Lantier L.AMPK in skeletal muscle function and metabo-
lism. FASEB J 32: 17411777, 2018. doi:10.1096/fj.201700442R.
555. McConell GK.Its well and truly time to stop stating that AMPK regu-
lates glucose uptake and fat oxidation during exercise. Am J
Physiol Endocrinol Metab 318: E564E567, 2020. doi:10.1152/
ajpendo.00511.2019.
556. Simcox J, Lamming DW.The central moTOR of metabolism. Dev
Cell 57: 691706, 2022. doi:10.1016/j.devcel.2022.02.024.
557. White JP.Amino acid trafcking and skeletal muscle protein synthe-
sis: a case of supply and demand. Front Cell Dev Biol 9: 656604,
2021. doi:10.3389/fcell.2021.656604.
558. Sanchez AM, Candau RB, Bernardi H.FoxO transcription factors:
their roles in the maintenance of skeletal muscle homeostasis. Cell
Mol Life Sci 71: 16571671, 2014. doi:10.1007/s00018-013-1513-z.
559. von Walden F.Ribosome biogenesis in skeletal muscle: coordina-
tion of transcription and translation. J Appl Physiol (1985) 127: 591
598, 2019. doi:10.1152/japplphysiol.00963.2018.
560. Vainshtein A, Sandri M.Signaling pathways that control muscle
mass. Int J Mol Sci 21: 4759, 2020. doi:10.3390/ijms21134759.
561. Yoshida T, Delafontaine P.Mechanisms of IGF-1-mediated regula-
tion of skeletal muscle hypertrophy and atrophy. Cells 9: 1970,
2020. doi:10.3390/cells9091970.
562. Markby GR, Sakamoto K.Transcription factor EB and TFE3: new
metabolic coordinators mediating adaptive responses to exercise
in skeletal muscle? Am J Physiol Endocrinol Metab 319: E763
E768, 2020. doi:10.1152/ajpendo.00339.2020.
563. Mesquita PH, Vann CG, Phillips SM, McKendry J, Young KC, Kavazis
AN, Roberts MD.Skeletal muscle ribosome and mitochondrial bio-
genesis in response to different exercise training modalities. Front
Physiol 12: 725866, 2021. doi:10.3389/fphys.2021.725866.
564. Atherton PJ, Babraj J, Smith K, Singh J, Rennie MJ, Wackerhage H.
Selective activation of AMPK-PGC-1alpha or PKB-TSC2-mTOR sig-
naling can explain specic adaptive responses to endurance or re-
sistance training-like electrical muscle stimulation. FASEB J 19:
786788, 2005. doi:10.1096/fj.04-2179fje.
565. Coffey VG, Zhong Z, Shield A, Canny BJ, Chibalin AV, Zierath JR,
Hawley JA.Early signaling responses to divergent exercise stimuli
in skeletal muscle from well-trained humans. FASEB J 20: 190192,
2006. doi:10.1096/fj.05-4809fje.
566. Cant
o C, Jiang LQ, Deshmukh AS, Mataki C, Coste A, Lagouge M,
Zierath JR, Auwerx J.Interdependence of AMPK and SIRT1 for
metabolic adaptation to fasting and exercise in skeletal muscle.
Cell Metab 11: 213219, 2010. doi:10.1016/j.cmet.2010.02.006.
567. Gurd BJ.Deacetylation of PGC-1alpha by SIRT1: importance for skele-
tal muscle function and exercise-induced mitochondrial biogenesis.
Appl Physiol Nutr Metab 36: 589597, 2011. doi:10.1139/h11-070.
568. DominyJE Jr, Lee Y, Gerhart-Hines Z, Puigserver P.Nutrient-depend-
ent regulation of PGC-1alphas acetylation state and metabolic func-
tion through the enzymatic activities of Sirt1/GCN5. Biochim Biophys
Acta 1804: 16761683, 2010. doi:10.1016/j.bbapap.2009.11.023.
569. Radak Z, Suzuki K, Posa A, Petrovszky Z, Koltai E, Boldogh I.The
systemic role of SIRT1 in exercise mediated adaptation. Redox Biol
35: 101467, 2020. doi:10.1016/j.redox.2020.101467.
570. Vargas-Ortiz K, P
erez-Vázquez V, Macias-Cervantes MH.Exercise
and sirtuins: a way to mitochondrial health in skeletal muscle. Int J
Mol Sci 20: 2717, 2019. doi:10.3390/ijms20112717.
571. Ramazi S, Zahiri J.Posttranslational modications in proteins:
resources, tools and prediction methods. Database (Oxford) 2021:
baab012, 2021. doi:10.1093/database/baab012.
572. Conibear AC.Deciphering protein post-translational modications
using chemical biology tools. Nat Rev Chem 4: 674695, 2020.
doi:10.1038/s41570-020-00223-8.
573. Memme JM, Hood DA.Molecular basis for the therapeutic effects
of exercise on mitochondrial defects. Front Physiol 11: 615038,
2020. doi:10.3389/fphys.2020.615038.
574. Popov DV.Adaptation of skeletal muscles to contractile activity of
varying duration and intensity: the role of PGC-1alpha. Biochemistry
(Mosc) 83: 613628, 2018. doi:10.1134/S0006297918060019.
575. Svensson K, Albert V, Cardel B, Salatino S, Handschin C.Skeletal
muscle PGC-1alpha modulates systemic ketone body homeostasis
and ameliorates diabetic hyperketonemia in mice. FASEB J 30:
19761986, 2016. doi:10.1096/fj.201500128.
576. Chan MC, Arany Z.The many roles of PGC-1alpha in muscle
recent developments. Metabolism 63: 441451, 2014. doi:10.1016/j.
metabol.2014.01.006.
577. Summermatter S, Santos G, P
erez-Schindler J, Handschin C.
Skeletal muscle PGC-1alpha controls whole-body lactate homeosta-
sis through estrogen-related receptor alpha-dependent activation
of LDH B and repression of LDH A. Proc Natl Acad Sci USA 110:
87388743, 2013. doi:10.1073/pnas.1212976110.
578. Kr
amer AI, Handschin C.How epigenetic modications drive the
expression and mediate the action of PGC-1alpha in the regulation
of metabolism. Int J Mol Sci 20: 5449, 2019. doi:10.3390/
ijms20215449.
579. Janzen NR, Whiteld J, Hoffman NJ.Interactive roles for AMPK and
glycogen from cellular energy sensing to exercise metabolism. Int J
Mol Sci 19: 3344, 2018. doi:10.3390/ijms19113344.
580. Richards P, Ourabah S, Montagne J, Burnol AF, Postic C, Guilmeau
S.MondoA/ChREBP: the usual suspects of transcriptional glucose
sensing; implication in pathophysiology. Metabolism 70: 133151,
2017. doi:10.1016/j.metabol.2017.01.033.
581. Goodman CA.Role of mTORC1 in mechanically induced increases
in translation and skeletal muscle mass. J Appl Physiol (1985) 127:
581590, 2019. doi:10.1152/japplphysiol.01011.2018.
582. Valenzuela PL, Morales JS, Emanuele E, Pareja-Galeano H, Lucia A.
Supplements with purported effects on muscle mass and strength.
Eur J Nutr 58: 29833008, 2019. doi:10.1007/s00394-018-1882-z.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1777
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
583. Reddy A, Bozi LH, Yaghi OK, Mills EL, Xiao H, Nicholson HE,
Paschini M, Paulo JA, Garrity R, Laznik-Bogoslavski D, Ferreira JC,
Carl CS, Sjøberg KA, Wojtaszewski JF, Jeppesen JF, Kiens B, Gygi
SP, Richter EA, Mathis D, Chouchani ET.pH-Gated succinate secre-
tion regulates muscle remodeling in response to exercise. Cell 183:
6275.e17, 2020. doi:10.1016/j.cell.2020.08.039.
584. Haws SA, Leech CM, Denu JM.Metabolism and the epigenome: a
dynamic relationship. Trends Biochem Sci 45: 731747, 2020.
doi:10.1016/j.tibs.2020.04.002.
585. McGee SL, Hargreaves M.Epigenetics and exercise. Trends
Endocrinol Metab 30: 636645, 2019. doi:10.1016/j.tem.2019.06.002.
586. Belhaj MR, Lawler NG, Hoffman NJ.Metabolomics and lipidomics:
expanding the molecular landscape of exercise biology. Metabolites
11: 151, 2021. doi:10.3390/metabo11030151.
587. Dankel SJ, Mattocks KT, Jessee MB, Buckner SL, Mouser JG,
Loenneke JP.Do metabolites that are produced during resistance
exercise enhance muscle hypertrophy? Eur J Appl Physiol 117:
21252135, 2017. doi:10.1007/s00421-017-3690-1.
588. Ibrahim A, Neinast M, Arany ZP.Myobolites: muscle-derived metab-
olites with paracrine and systemic effects. Curr Opin Pharmacol
34: 1520, 2017. doi:10.1016/j.coph.2017.03.007.
589. Maurer J, Hoene M, Weigert C.Signals from the circle: tricarboxylic
acid cycle intermediates as myometabokines. Metabolites 11: 474,
2021. doi:10.3390/metabo11080474.
590. Murphy RM, Watt MJ, Febbraio MA.Metabolic communication dur-
ing exercise. Nat Metab 2: 805816, 2020. doi:10.1038/s42255-
020-0258-x.
591. Agudelo LZ, Femenía T, Orhan F, Porsmyr-Palmertz M, Goiny M,
Martinez-Redondo V, Correia JC, Izadi M, Bhat M, Schuppe-
Koistinen I, Pettersson AT, Ferreira DMS, Krook A, Barres R, Zierath
JR, Erhardt S, Lindskog M, Ruas JL.Skeletal muscle PGC-1alpha1
modulates kynurenine metabolism and mediates resilience to
stress-induced depression. Cell 159: 3345, 2014. doi:10.1016/j.
cell.2014.07.051.
592. Wojtaszewski JF, MacDonald C, Nielsen JN, Hellsten Y, Hardie DG,
Kemp BE, Kiens B, Richter EA.Regulation of 50AMP-activated pro-
tein kinase activity and substrate utilization in exercising human
skeletal muscle. Am J Physiol Endocrinol Metab 284: E813E822,
2003. doi:10.1152/ajpendo.00436.2002.
593. Lindholm ME, Rundqvist H.Skeletal muscle hypoxia-inducible fac-
tor-1 and exercise. Exp Physiol 101: 2832, 2016. doi:10.1113/
EP085318.
594. Brugniaux JV, Coombs GB, Barak OF, Dujic Z, Sekhon MS, Ainslie
PN.Highs and lows of hyperoxia: physiological, performance, and
clinical aspects. Am J Physiol Regul Integr Comp Physiol 315: R1
R27, 2018. doi:10.1152/ajpregu.00165.2017.
595. Poole DC, Pittman RN, Musch TI, Østergaard L.August Kroghs
theory of muscle microvascular control and oxygen delivery: a para-
digm shift based on new data. JPhysiol598: 44734507, 2020.
doi:10.1113/JP279223.
596. Clanton TL.Managing the power grid: how myoglobin can regulate
PO
2
and energy distribution in skeletal muscle. J Appl Physiol
(1985) 126: 787790, 2019. doi:10.1152/japplphysiol.00614.2018.
597. Richardson RS, Newcomer SC, Noyszewski EA.Skeletal muscle in-
tracellular PO
2
assessed by myoglobin desaturation: response to
graded exercise. J Appl Physiol (1985) 91: 26792685, 2001.
doi:10.1152/jappl.2001.91.6.2679.
598. Bock JM, Kruse NT, Donnelly C, Hirai DM, Craig JC, Colburn TD,
Musch TI, Poole DC, Rosenberry R, Tian F, Liu H, Nelson MD,
Piknova B, Willis WT, Zuo L, Zhou T, Riveros-Rivera A, Cristancho E,
Gunga HC.Commentaries on Viewpoint: Managing the power grid:
how myoglobin can regulate Po
2
and energy distribution in skeletal
muscle. J Appl Physiol (1985) 126: 791794, 2019. doi:10.1152/
japplphysiol.01107.2018.
599. Brugarolas J, Lei K, Hurley RL, Manning BD, Reiling JH, Hafen E,
Witters LA, Ellisen LW, Kaelin WG Jr. Regulation of mTOR function
in response to hypoxia by REDD1 and the TSC1/TSC2 tumor sup-
pressor complex. Genes Dev 18: 28932904, 2004. doi:10.1101/
gad.1256804.
600. Rasbach KA, Gupta RK, Ruas JL, Wu J, Naseri E, Estall JL,
Spiegelman BM.PGC-1alpha regulates a HIF2alpha-dependent
switch in skeletal muscle ber types. Proc Natl Acad Sci USA 107:
2186621871, 2010. doi:10.1073/pnas.1016089107.
601. Conceição MS, Chacon-Mikahil MP, Telles GD, Libardi CA, Júnior
EM, Vechin FC, De Andrade AL, Gáspari AF, Brum PC, Cavaglieri
CR, Serag S, Spiegelman BM, Hawley JA, Camera DM.Attenuated
PGC-1alpha isoforms following endurance exercise with blood ow
restriction. Med Sci Sports Exerc 48: 16991707, 2016. doi:10.1249/
MSS.0000000000000970.
602. Barclay CJ.The basis of differences in thermodynamic efciency
among skeletal muscles. Clin Exp Pharmacol Physiol 44: 1279
1286, 2017. doi:10.1111/1440-1681.12850.
603. Archer AE, Von Schulze AT, Geiger PC.Exercise, heat shock pro-
teins and insulin resistance. Philos Trans R Soc Lond B Biol Sci
373: 20160529, 2018. doi:10.1098/rstb.2016.0529.
604. Henstridge DC, Febbraio MA, Hargreaves M.Heat shock proteins
and exercise adaptations. Our knowledge thus far and the road still
ahead. J Appl Physiol (1985) 120: 683691, 2016. doi:10.1152/
japplphysiol.00811.2015.
605. McGorm H, Roberts LA, Coombes JS, Peake JM.Turning up the
heat: an evaluation of the evidence for heating to promote exercise
recovery, muscle rehabilitation and adaptation. Sports Med 48:
13111328, 2018. doi:10.1007/s40279-018-0876-6.
606. Petersen AC, Fyfe JJ.Post-exercise cold water immersion effects
on physiological adaptations to resistance training and the underly-
ing mechanisms in skeletal muscle: a narrative review. Front Sports
Act Living 3: 660291, 2021. doi:10.3389/fspor.2021.660291.
607. Dubois B, Esculier JF.Soft-tissue injuries simply need PEACE and
LOVE. Br J Sports Med 54: 7273, 2020. doi:10.1136/bjsports-2019-
101253.
608. Bohnert KR, McMillan JD, Kumar A.Emerging roles of ER stress and
unfolded protein response pathways in skeletal muscle health and
disease. J Cell Physiol 233: 6778, 2018. doi:10.1002/jcp.25852.
609. Marafon BB, Pinto AP, Ropelle ER, de Moura LP, Cintra DE, Pauli JR,
da Silva AS.Muscle endoplasmic reticulum stress in exercise. Acta
Physiol (Oxf) 235: e13799, 2022. doi:10.1111/apha.13799.
610. Zhang Y, Oliveira AN, Hood DA.The intersection of exercise and
aging on mitochondrial protein quality control. Exp Gerontol 131:
110824, 2020. doi:10.1016/j.exger.2019.110824.
611. Song J, Herrmann JM, Becker T.Quality control of the mitochon-
drial proteome. Nat Rev Mol Cell Biol 22: 5470, 2021. doi:10.1038/
s41580-020-00300-2.
612. Pickles S, Vigi
eP,YouleRJ.Mitophagy and quality control mecha-
nisms in mitochondrial maintenance. Curr Biol 28: R170R185,
2018. doi:10.1016/j.cub.2018.01.004.
FURRER ET AL.
1778 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
613. Soubannier V, McLelland GL, Zunino R, Braschi E, Rippstein P, Fon
EA, McBride HM.A vesicular transport pathway shuttles cargo from
mitochondria to lysosomes. Curr Biol 22: 135141, 2012. doi:10.1016/j.
cub.2011.11.057.
614. Sanchez AM, Candau R, Bernardi H.Recent data on cellular compo-
nent turnover: focus on adaptations to physical exercise. Cells 8:
542, 2019. doi:10.3390/cells8060542.
615. Romanello V, Sandri M.Implications of mitochondrial fusion and s-
sion in skeletal muscle mass and health. Semin Cell Dev Biol 2022:
S1084-9521(22)00050-7, 2022. doi:10.1016/j.semcdb.2022.02.011.
616. Tanaka T, Nishimura A, Nishiyama K, Goto T, Numaga-Tomita T,
Nishida M.Mitochondrial dynamics in exercise physiology. Pugers
Arch 472: 137153, 2020. doi:10.1007/s00424-019-02258-3.
617. Hood DA, Memme JM, Oliveira AN, Triolo M.Maintenance of skele-
tal muscle mitochondria in health, exercise, and aging. Annu Rev
Physiol 81: 1941, 2019. doi:10.1146/annurev-physiol-020518-114310.
618. Guan Y, Drake JC, Yan Z.Exercise-induced mitophagy in skeletal
muscle and heart. Exerc Sport Sci Rev 47: 151156, 2019.
doi:10.1249/JES.0000000000000192.
619. Ament W, Verkerke GJ.Exercise and fatigue. Sports Med 39: 389
422, 2009. doi:10.2165/00007256-200939050-00005.
620. Meeusen R, Van Cutsem J, Roelands B.Endurance exercise-
induced and mental fatigue and the brain. Exp Physiol 106: 2294
2298, 2021. doi:10.1113/EP088186.
621. Taylor JL, Amann M, Duchateau J, Meeusen R, Rice CL.Neural con-
tributions to muscle fatigue: from the brain to the muscle and back
again. Med Sci Sports Exerc 48: 22942306, 2016. doi:10.1249/
MSS.0000000000000923.
622. Proschinger S, Freese J.Neuroimmunological and neuroenergetic
aspects in exercise-induced fatigue. Exerc Immunol Rev 25: 819,
2019.
623. Noakes TD.Fatigue is a brain-derived emotion that regulates the
exercise behavior to ensure the protection of whole body homeo-
stasis. Front Physiol 3: 82, 2012. doi:10.3389/fphys.2012.00082.
624. Bawa P.Neural control of motor output: can training change it?
Exerc Sport Sci Rev 30: 5963, 2002. doi:10.1097/00003677-
200204000-00003.
625. Heckman CJ, Enoka RM.Motor unit. Compr Physiol 2: 26292682,
2012. doi:10.1002/cphy.c100087.
626. Romer LM, Polkey MI.Exercise-induced respiratory muscle fatigue:
implications for performance. J Appl Physiol (1985) 104: 879888,
2008. doi:10.1152/japplphysiol.01157.2007.
627. Thomas K, Goodall S, Howatson G.Performance fatigability is not
regulated to a peripheral critical threshold. Exerc Sport Sci Rev 46:
240246, 2018. doi:10.1249/JES.0000000000000162.
628. Finsterer J.Biomarkers of peripheral muscle fatigue during exer-
cise. BMC Musculoskelet Disord 13: 218, 2012. doi:10.1186/1471-
2474-13-218.
629. Cheng AJ, Place N, Westerblad H.molecular basis for exercise-
induced fatigue: the importance of strictly controlled cellular Ca
21
handling. Cold Spring Harb Perspect Med 8: a029710, 2018.
doi:10.1101/cshperspect.a029710.
630. Powers SK, Schrager M.Redox signaling regulates skeletal muscle
remodeling in response to exercise and prolonged inactivity.
Redox Biol 54: 102374, 2022. doi:10.1016/j.redox.2022.102374.
631. McKeegan K, Mason SA, Trewin AJ, Keske MA, Wadley GD, Della
Gatta PA, Nikolaidis MG, Parker L.Reactive oxygen species in exer-
cise and insulin resistance: working towards personalized antioxi-
dant treatment. Redox Biol 44: 102005, 2021. doi:10.1016/j.
redox.2021.102005.
632. Giordani L, He GJ, Negroni E, Sakai H, Law JY, Siu MM, Wan R,
Corneau A, Tajbakhsh S, Cheung TH, Le Grand F.High-dimensional
single-cell cartography reveals novel skeletal muscle-resident cell
populations. Mol Cell 74: 609621.e6, 2019. doi:10.1016/j.molcel.
2019.02.026.
633. Rubenstein AB, Smith GR, Raue U, Begue G, Minchev K, Ruf-
Zamojski F, Nair VD, Wang X, Zhou L, Zaslavsky E, Trappe TA,
Trappe S, Sealfon SC.Single-cell transcriptional proles in human
skeletal muscle. Sci Rep 10: 229, 2020. doi:10.1038/s41598-019-
57110-6.
634. De Micheli AJ, Laurilliard EJ, Heinke CL, Ravichandran H, Fraczek P,
Soueid-Baumgarten S, De Vlaminck I, Elemento O, Cosgrove BD.
Single-cell analysis of the muscle stem cell hierarchy identies het-
erotypic communication signals involved in skeletal muscle regen-
eration. Cell Rep 30: 35833595.e5, 2020. doi:10.1016/j.celrep.
2020.02.067.
635. Morgan J, Partridge T.Skeletal muscle in health and disease. Dis
Model Mech 13: dmm042192, 2020. doi:10.1242/dmm.042192.
636. Forcina L, Cosentino M, Musar
oA.Mechanisms regulating muscle
regeneration: insights into the interrelated and time-dependent phases
of tissue healing. Cells 9: 1297, 2020. doi:10.3390/cells9051297.
637. Peake JM, Neubauer O, Della Gatta PA, Nosaka K.Muscle damage
and inammation during recovery from exercise. J Appl Physiol
(1985) 122: 559570, 2017. doi:10.1152/japplphysiol.00971.2016.
638. Markus I, Constantini K, Hoffman JR, Bartolomei S, Gepner Y.
Exercise-induced muscle damage: mechanism, assessment and
nutritional factors to accelerate recovery. Eur J Appl Physiol 121:
969992, 2021. doi:10.1007/s00421-020-04566-4.
639. Ziemkiewicz N, Hilliard G, Pullen NA, Garg K.The role of innate and
adaptive immune cells in skeletal muscle regeneration. Int J Mol
Sci 22: 3265, 2021. doi:10.3390/ijms22063265.
640. Burzyn D, Kuswanto W, Kolodin D, Shadrach JL, Cerletti M, Jang Y,
Sek E, Tan TG, Wagers AJ, Benoist C, Mathis D.A special popula-
tion of regulatory T cells potentiates muscle repair. Cell 155: 1282
1295, 2013. doi:10.1016/j.cell.2013.10.054.
641. Paulsen G, Crameri R, Benestad HB, Fjeld JG, Morkrid L, Hall
en J,
Raastad T.Time course of leukocyte accumulation in human mus-
cle after eccentric exercise. Med Sci Sports Exerc 42: 7585, 2010.
doi:10.1249/MSS.0b013e3181ac7adb.
642. Bengtsen M, Winje IM, Eftestøl E, Landskron J, Sun C, Nygård K,
Domanska D, Millay DP, Meza-Zepeda LA, Gundersen K.Comparing
the epigenetic landscape in myonuclei puried with a PCM1 antibody
from a fast/glycolytic and a slow/oxidative muscle. PLoS Genet 17:
e1009907, 2021. doi:10.1371/journal.pgen.1009907.
643. Dinulovic I, Furrer R, Di Fulvio S, Ferry A, Beer M, Handschin C.
PGC-1alpha modulates necrosis, inammatory response, and
brotic tissue formation in injured skeletal muscle. Skelet Muscle 6:
38, 2016. doi:10.1186/s13395-016-0110-x.
644. Handschin C, Mortezavi A, Plock J, Eberli D.External physical and
biochemical stimulation to enhance skeletal muscle bioengineer-
ing. Adv Drug Deliv Rev 82-83: 168175, 2015. doi:10.1016/j.
addr.2014.10.021.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1779
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
645. Roman W, Mu~
noz-Cánoves P.Muscle is a stage, and cells and fac-
tors are merely players. Trends Cell Biol 32: 835840, 2022.
doi:10.1016/j.tcb.2022.03.001.
646. Barth
el
emyF,DefourA,L
evy N, Krahn M, Bartoli M.Muscle cells x
breaches by orchestrating a membrane repair ballet. JNeuromuscul
Dis 5: 2128, 2018. doi:10.3233/JND-170251.
647. Vassilopoulos S.Unconventional roles for membrane trafcpro-
teins in response to muscle membrane stress. Curr Opin Cell Biol
65: 4249, 2020. doi:10.1016/j.ceb.2020.02.007.
648. WhitsonBA,TanT,GongN,ZhuH,MaJ.Muscle multiorgan cross-
talk with MG53 as a myokine for tissue repair and regeneration. Curr
Opin Pharmacol 59: 2632, 2021. doi:10.1016/j.coph.2021.04.005.
649. Roman W, Pinheiro H, Pimentel MR, Segal
es J, Oliveira LM, García-
Domínguez E, G
omez-Cabrera MC, Serrano AL, Gomes ER, Munoz-
Canoves P.Muscle repair after physiological damage relies on nu-
clear migration for cellular reconstruction. Science 374: 355359,
2021. doi:10.1126/science.abe5620.
650. Taylor-Weiner H, Grigsby CL, Ferreira DM, Dias JM, Stevens MM,
Ruas JL, Teixeira AI.Modeling the transport of nuclear proteins
along single skeletal muscle cells. Proc Natl Acad Sci USA 117:
29782986, 2020. doi:10.1073/pnas.1919600117.
651. Denes LT, Kelley CP, Wang ET.Microtubule-based transport is
essential to distribute RNA and nascent protein in skeletal muscle.
Nat Commun 12: 6079, 2021. doi:10.1038/s41467-021-26383-9.
652. Hotel T, Freiwald J, Hoppe MW, Lutter C, Forst R, Grim C, Bloch
W, H
uttel M, Heiss R.Advances in delayed-onset muscle sore-
ness (DOMS): Part I: pathogenesis and diagnostics. Sportverletz
Sportschaden 32: 243250, 2018. doi:10.1055/a-0753-1884.
653. Urso ML.Anti-inammatory interventions and skeletal muscle injury:
benetordetriment?J Appl Physiol (1985) 115: 920928, 2013.
doi:10.1152/japplphysiol.00036.2013.
654. Heiss R, Lutter C, Freiwald J, Hoppe MW, Grim C, Poettgen K, Forst
R, Bloch W, H
uttel M, Hotel T.Advances in delayed-onset muscle
soreness (DOMS): Part II: treatment and prevention. Sportverletz
Sportschaden 33: 2129, 2019. doi:10.1055/a-0810-3516.
655. Deyhle MR, Gier AM, Evans KC, Eggett DL, Nelson WB, Parcell AC,
Hyldahl RD.Skeletal muscle inammation following repeated bouts
of lengthening contractions in humans. Front Physiol 6: 424, 2015.
doi:10.3389/fphys.2015.00424.
656. Damas F, Phillips SM, Libardi CA, Vechin FC, Lixandrão ME, Jannig
PR, Costa LA, Bacurau AV, Snijders T, Parise G,Tricoli V, Roschel H,
Ugrinowitsch C.Resistance training-induced changes in integrated
myobrillar protein synthesis are related to hypertrophy only after
attenuation of muscle damage. J Physiol 594: 52095222, 2016.
doi:10.1113/JP272472.
657. Weakley J, Halson SL, Mujika I.Overtraining syndrome symptoms
and diagnosis in athletes: where is the research? A systematic
review. Int J Sports Physiol Perform 17: 675681, 2022. doi:10.1123/
ijspp.2021-0448.
658. Solsona R, Pavlin L, Bernardi H, Sanchez AM.Molecular regulation
of skeletal muscle growth and organelle biosynthesis: practical rec-
ommendations for exercise training. Int J Mol Sci 22: 2741, 2021.
doi:10.3390/ijms22052741.
659. Koves TR, Sparks LM, Kovalik JP, Mosedale M, Arumugam R,
DeBalsi KL, Everingham K, Thorne L, Phielix E, Meex RC, Kien CL,
Hesselink MK, Schrauwen P, Muoio DM.PPARgamma coactivator-
1alpha contributes to exercise-induced regulation of intramuscular
lipid droplet programming in mice and humans. J Lipid Res 54:
522534, 2013. doi:10.1194/jlr.P028910.
660. Gemmink A, Schrauwen P, Hesselink MK.Exercising your fat (me-
tabolism) into shape: a muscle-centred view. Diabetologia 63:
14531463, 2020. doi:10.1007/s00125-020-05170-z.
661. Koronowski KB, Sassone-Corsi P.Communicating clocks shape cir-
cadian homeostasis. Science 371: eabd0951, 2021. doi:10.1126/
science.abd0951.
662. Gabriel BM, Zierath JR.Circadian rhythms and exercise - re-setting
the clock in metabolic disease. Nat Rev Endocrinol 15: 197206,
2019. doi:10.1038/s41574-018-0150-x.
663. Gutierrez-Monreal MA, Harmsen JF, Schrauwen P, Esser KA.
Ticking for metabolic health: the skeletal-muscle clocks. Obesity
(Silver Spring) 28, Suppl 1: S46S54, 2020. doi:10.1002/oby.22826.
664. Gabriel BM, Zierath JR.Zeitgebers of skeletal muscle and implica-
tions for metabolic health. J Physiol 600: 10271036, 2022.
doi:10.1113/JP280884.
665. Healy KL, Morris AR, Liu AC.Circadian synchrony: sleep, nutrition, and
physical activity. Front Netw Physiol 1: 732243, 2021. doi:10.3389/
fnetp.2021.732243.
666. Mansingh S, Handschin C.Time to train: the involvement of the mo-
lecular clock in exercise adaptation of skeletal muscle. Front
Physiol 13: 902031, 2022. doi:10.3389/fphys.2022.902031.
667. Morris AR, Stanton DL, Roman D, Liu AC.Systems level understand-
ing of circadian integration with cell physiology. J Mol Biol 432:
35473564, 2020. doi:10.1016/j.jmb.2020.02.002.
668. Maier G, Delezie J, Westermark PO, Santos G, Ritz D, Handschin C.
Transcriptomic, proteomic and phosphoproteomic underpinnings
of daily exercise performance and zeitgeber activity of training in
mouse muscle. J Physiol 600: 769796, 2022. doi:10.1113/
JP281535.
669. Graham ZA, Lavin KM, OBryan SM, Thalacker-Mercer AE, Buford
TW, Ford KM, Broderick TJ, Bamman MM.Mechanisms of exercise
as a preventative measure to muscle wasting. Am J Physiol Cell
Physiol 321: C40C57, 2021. doi:10.1152/ajpcell.00056.2021.
670. Harper C, Gopalan V, Goh J.Exercise rescues mitochondrial cou-
pling in aged skeletal muscle: a comparison of different modalities
in preventing sarcopenia. J Transl Med 19: 71, 2021. doi:10.1186/
s12967-021-02737-1.
671. Larsson L, Degens H, Li M, Salviati L, Lee YI, Thompson W, Kirkland
JL, Sandri M.Sarcopenia: aging-related loss of muscle mass and
function. Physiol Rev 99: 427511, 2019. doi:10.1152/physrev.
00061.2017.
672. Leal LG, Lopes MA, Peres SB, Batista ML Jr. Exercise training as
therapeutic approach in cancer cachexia: a review of potential anti-
inammatory effect on muscle wasting. Front Physiol 11: 570170,
2020. doi:10.3389/fphys.2020.570170.
673. Deane CS, Willis CR, Phillips BE, AthertonPJ, Harries LW, AmesRM,
Szewczyk NJ, Etheridge T.Transcriptomic meta-analysis of disuse
muscle atrophy vs. resistance exercise-induced hypertrophy in
young and older humans. J Cachexia Sarcopenia Muscle 12: 629
645, 2021. doi:10.1002/jcsm.12706.
674. Domin R, Dadej D, Pytka M, Zybek-Kocik A, RuchałaM,GuzikP.
Effect of various exercise regimens on selected exercise-induced
cytokines in healthy people. Int J Environ Res Public Health 18:
1261, 2021. doi:10.3390/ijerph18031261.
FURRER ET AL.
1780 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
675. Petrosino JM, Hinger SA, Golubeva VA, Barajas JM, Dorn LE, Iyer
CC, Sun HL, Arnold WD, He C, Accornero F.The m(6)A methyltrans-
ferase METTL3 regulates muscle maintenance and growth in mice.
Nat Commun 13: 168, 2022. doi:10.1038/s41467-021-27848-7.
676. Duan K, Gao X, Zhu D.The clinical relevance and mechanism of
skeletal muscle wasting. Clin Nutr 40: 2737, 2021. doi:10.1016/j.
clnu.2020.07.029.
677. Furrer R, Handschin C.Muscle wasting diseases: novel targets and
treatments. Annu Rev Pharmacol Toxicol 59: 315339, 2019.
doi:10.1146/annurev-pharmtox-010818-021041.
678. Sartori R, Romanello V, Sandri M.Mechanisms of muscle atrophy
and hypertrophy: implications in health and disease. Nat Commun
12: 330, 2021. doi:10.1038/s41467-020-20123-1.
679. Ebert SM, Al-Zougbi A, Bodine SC, Adams CM.Skeletal muscle atro-
phy: discovery of mechanisms and potential therapies. Physiology
(Bethesda) 34: 232239, 2019. doi:10.1152/physiol.00003.2019.
680. Verbrugge SA, Sch
onfelder M, Becker L, Yaghoob Nezhad F,
Hrab
e de Angelis M, Wackerhage H.Genes whose gain or loss-of-
function increases skeletal muscle mass in mice: a systematic litera-
ture review. Front Physiol 9: 553, 2018. doi:10.3389/fphys.2018.
00553.
681. Yaghoob Nezhad F, Verbrugge SA, Sch
onfelder M, Becker L, Hrab
e
de Angelis M, Wackerhage H.Genes whose gain or loss-of-function
increases endurance performance in mice: a systematic literature
review. Front Physiol 10: 262, 2019. doi:10.3389/fphys.2019.00262.
682. Qi Z, Zhai X, Ding S.How to explain exercise-induced phenotype
from molecular data: rethink and reconstruction based on AMPK
and mTOR signaling. Springerplus 2: 693, 2013. doi:10.1186/2193-
1801-2-693.
683. Hingst JR, Kjøbsted R, Birk JB, Jørgensen NO, Larsen MR, Kido K,
Larsen JK, Kjeldsen SA, Fentz J, Frøsig C, Holm S, Fritzen AM,
Dohlmann TL, Larsen S, Foretz M, Viollet B, Schjerling P, Overby P,
Halling JF, Pilegaard H, Hellsten Y, Wojtaszewski JF.Inducible dele-
tion of skeletal muscle AMPKalpha reveals that AMPK is required
for nucleotide balance but dispensable for muscle glucose uptake
and fat oxidation during exercise. Mol Metab 40: 101028, 2020.
doi:10.1016/j.molmet.2020.101028.
684. You JS, McNally RM, Jacobs BL, Privett RE, Gundermann DM, Lin
KH, Steinert ND, Goodman CA, Hornberger TA.The role of raptor
in the mechanical load-induced regulation of mTOR signaling, pro-
tein synthesis, and skeletal muscle hypertrophy. FASEB J 33:
40214034, 2019. doi:10.1096/fj.201801653RR.
685. Castets P, Lin S, Rion N, Di Fulvio S, Romanino K, Guridi M, Frank S,
Tintignac LA, Sinnreich M, R
uegg MA.Sustained activation of
mTORC1 in skeletal muscle inhibits constitutive and starvation-
induced autophagy and causes a severe, late-onset myopathy. Cell
Metab 17: 731744, 2013. doi:10.1016/j.cmet.2013.03.015.
686. Orchard S, Ammari M, Aranda B, Breuza L, Briganti L, Broackes-
Carter F, et al.The MIntAct projectIntAct as a common curation
platform for 11 molecular interaction databases. Nucleic Acids Res
42: D358D363, 2014. doi:10.1093/nar/gkt1115.
687. Luciano AK, Korobkina ED, Lyons SP, Haley JA, Fluharty SM, Jung
SM, Kettenbach AN, Guertin DA.Proximity labeling of endogenous
RICTOR identies mTOR Complex 2 regulation by ADP ribosylation
factor ARF1. JBiolChem298: 102379, 2022. doi:10.1016/j.
jbc.2022.102379.
688. Alves MM, Fuhler GM, Queiroz KC, Scholma J, Goorden S, Anink J,
Spek CA, Hoogeveen-Westerveld M, Bruno MJ, Nellist M, Elgersma
Y, Aronica E, Peppelenbosch MP.PAK2 is an effector of TSC1/2
signaling independent of mTOR and a potential therapeutic target
for Tuberous Sclerosis Complex. Sci Rep 5: 14534, 2015.
doi:10.1038/srep14534.
689. Sun L, Yan Y, Lv H, Li J, Wang Z, Wang K, Wang L, Li Y, Jiang H,
Zhang Y.Rapamycin targets STAT3 and impacts c-Myc to suppress
tumor growth. Cell Chem Biol 29: 373385.e6, 2022. doi:10.1016/j.
chembiol.2021.10.006.
690. Guadagnin E, Mázala D, Chen YW.STAT3 in skeletal muscle func-
tion and disorders. Int J Mol Sci 19: 2265, 2018. doi:10.3390/
ijms19082265.
691. Seaborne RA, Sharples AP.The interplay between exercise metabo-
lism, epigenetics, and skeletal muscle remodeling. Exerc Sport Sci
Rev 48: 188200, 2020. doi:10.1249/JES.0000000000000227.
692. Tucker R, Collins M.What makes champions? A review of the rela-
tive contribution of genes and training to sporting success. Br J
Sports Med 46: 555561, 2012. doi:10.1136/bjsports-2011-090548.
693. Yan X, Papadimitriou I, Lidor R, Eynon N.Nature versus nurture in
determining athletic ability. Med Sport Sci 61: 1528, 2016.
doi:10.1159/000445238.
694. Eynon N, Hanson ED, Lucia A, Houweling PJ, Garton F, North KN,
Bishop DJ.Genes for elite power and sprint performance: ACTN3
leads the way. Sports Med 43: 803817, 2013. doi:10.1007/s40279-
013-0059-4.
695. Papadimitriou ID, Lucia A, Pitsiladis YP, Pushkarev VP, Dyatlov DA,
Orekhov EF, Artioli GG, Guilherme JP, Lancha AH Jr, Ginevi
cienėV,
Cieszczyk P, Maciejewska-Karlowska A, Sawczuk M, Muniesa CA,
Kouvatsi A, Massidda M, Cal
o CM, Garton F, Houweling PJ, Wang
G, Austin K, Druzhevskaya AM, Astratenkova IV, Ahmetov II, Bishop
DJ, North KN, Eynon N.ACTN3 R577X and ACE I/D gene variants
inuence performance in elite sprinters: a multi-cohort study. BMC
Genomics 17: 285, 2016. doi:10.1186/s12864-016-2462-3.
696. Rankinen T, Fuku N, Wolfarth B, Wang G, Sarzynski MA, Alexeev
DG, et al.No Evidence of a common DNA variant prole specicto
world class endurance athletes. PLoS One 11: e0147330, 2016.
doi:10.1371/journal.pone.0147330.
697. Bouchard C.Genomic predictors of trainability. Exp Physiol 97:
347352, 2012. doi:10.1113/expphysiol.2011.058735.
698. Bouchard C, Sarzynski MA, Rice TK, Kraus WE, Church TS, Sung YJ,
Rao DC, Rankinen T.Genomic predictors of the maximal O
2
uptake
response to standardized exercise training programs. JAppl
Physiol (1985) 110: 11601170, 2011. doi:10.1152/japplphysiol.00973.
2010.
699. Bouchard C, An P, Rice T, Skinner JS, Wilmore JH, Gagnon J,
P
erusse L, Leon AS, Rao DC.Familial aggregation of Vo
2max
response to exercise training: results from the HERITAGE Family
Study. J Appl Physiol (1985) 87: 10031008, 1999. doi:10.1152/
jappl.1999.87.3.1003.
700. Ruegsegger GN, Booth FW.Health benets of exercise. Cold Spring
Harb Perspect Med 8: a029694, 2018. doi:10.1101/cshperspect.
a029694.
701. Pickering C, Kiely J.Do non-responders to exercise exist-and if so,
what should we do about them? Sports Med 49: 17, 2019.
doi:10.1007/s40279-018-01041-1.
702. Montero D, Lundby C.Refuting the myth of non-response to exer-
cise training: non-respondersdo respond to higher dose of train-
ing. JPhysiol595: 33773387, 2017. doi:10.1113/JP273480.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1781
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
703. Booth FW, Laye MJ.The future: genes, physical activity and health.
Acta Physiol (Oxf) 199: 549556, 2010. doi:10.1111/j.1748-1716.2010.
02117.x.
704. Fiuza-Luces C, Garatachea N, Berger NA, Lucia A.Exercise is the
real polypill. Physiology (Bethesda) 28: 330358, 2013. doi:10.1152/
physiol.00019.2013.
705. Radak Z, Taylor AW.Issues on trainability. Front Physiol 12:
790196, 2021. doi:10.3389/fphys.2021.790196.
706. Koch LG, Pollott GE, Britton SL.Selectively bred rat model system
for low and high response to exercise training. Physiol Genomics
45: 606614, 2013. doi:10.1152/physiolgenomics.00021.2013.
707. Williams CJ, Li Z, Harvey N, Lea RA, Gurd BJ, Bonaglia JT, et al.
Genome wide association study of response to interval and contin-
uous exercise training: the Predict-HIIT study. JBiomedSci28: 37,
2021. doi:10.1186/s12929-021-00733-7.
708. Moir HJ, Kemp R, Folkerts D, Spendiff O, Pavlidis C, Opara E.Genes
and elite marathon running performance: a systematic review. J
Sports Sci Med 18: 559568, 2019.
709. Vernillo G, Schena F, Berardelli C, Rosa G, Galvani C, Maggioni M,
Agnello L, La Torre A.Anthropometric characteristics of top-class
Kenyan marathon runners. J Sports Med Phys Fitness 53: 403
408, 2013.
710. De Moor MH, Spector TD, Cherkas LF, Falchi M, Hottenga JJ,
Boomsma DI, De Geus EJ.Genome-wide linkage scan for athlete
status in 700 British female DZ twin pairs. Twin Res Hum Genet 10:
812820, 2007. doi:10.1375/twin.10.6.812.
711. Bouchard C, Daw EW, Rice T, P
erusse L, Gagnon J, Province MA,
LeonAS,RaoDC,SkinnerJS,WilmoreJH.Familial resemblance
for Vo
2max
in the sedentary state: the HERITAGE family study.
Med Sci Sports Exerc 30: 252258, 1998. doi:10.1097/
00005768-199802000-00013.
712. Konopka MJ, van den Bunder JC, Rietjens G, Sperlich B, Zeegers
MP.Genetics of long-distance runners and road cyclists-A system-
atic review with meta-analysis. Scand J Med Sci Sports 32: 1414
1429, 2022. doi:10.1111/sms.14212.
713. de la Chapelle A, Tr
askelin AL, Juvonen E.Truncated erythropoietin
receptor causes dominantly inherited benign human erythrocytosis.
Proc Natl Acad Sci USA 90: 44954499, 1993. doi:10.1073/
pnas.90.10.4495.
714. Ma F, Yang Y, Li X, Zhou F, Gao C, Li M, Gao L.The association of
sport performance with ACE and ACTN3 genetic polymorphisms: a
systematic review and meta-analysis. PLoS One 8: e54685, 2013.
doi:10.1371/journal.pone.0054685.
715. Montgomery HE, Marshall R, Hemingway H, Myerson S, Clarkson P,
Dollery C, Hayward M, Holliman DE, Jubb M, World M, Thomas EL,
Brynes AE, Saeed N, Barnard M, Bell JD, Prasad K, Rayson M,
Talmud PJ, Humphries SE.Human gene for physical performance.
Nature 393: 221222, 1998. doi:10.1038/30374.
716. Jacob Y, Spiteri T, Hart NH, Anderton RS.The potential role of
genetic markers in talent identication and athlete assessment in
elite sport. Sports (Basel) 6: 88, 2018. doi:10.3390/sports6030088.
717. Del Coso J, Hiam D, Houweling P, P
erez LM, Eynon N, Lucía A.
More than a speed gene: ACTN3 R577X genotype, trainability,
muscle damage, and the risk for injuries. Eur J Appl Physiol 119:
4960, 2019. doi:10.1007/s00421-018-4010-0.
718. Pickering C, Kiely J, Grgic J, Lucia A, Del Coso J.Can Genetic
Testing Identify Talent for Sport? Genes (Basel) 10: 972, 2019.
doi:10.3390/genes10120972.
719. Yang N, MacArthur DG, Gulbin JP, Hahn AG, Beggs AH, Easteal S,
North K.ACTN3 genotype is associated with human elite athletic
performance. Am J Hum Genet 73: 627631, 2003. doi:10.1086/
377590.
720. Yang R, Shen X, Wang Y, Voisin S, Cai G, Fu Y, Xu W, Eynon N,
Bishop DJ, Yan X.ACTN3 R577X gene variant is associated with
muscle-related phenotypes in elite Chinese sprint/power ath-
letes. JStrengthCondRes31: 11071115, 2017. doi:10.1519/JSC.
0000000000001558.
721. MacArthur DG, Seto JT, Raftery JM, Quinlan KG, Huttley GA, Hook
JW, Lemckert FA, Kee AJ, Edwards MR, Berman Y, Hardeman EC,
Gunning PW, Easteal S, Yang N, North KN.Loss of ACTN3 gene
function alters mouse muscle metabolism and shows evidence of
positive selection in humans. Nat Genet 39: 12611265, 2007.
doi:10.1038/ng2122.
722. Ahmetov II, Egorova ES, Gabdrakhmanova LJ, Fedotovskaya ON.
Genes and athletic performance: an update. Med Sport Sci 61: 41
54, 2016. doi:10.1159/000445240.
723. Bye A, Klevjer M, Ryeng E, Silva GJ, Moreira JB, Stensvold D,
Wisløff U.Identication of novel genetic variants associated with
cardiorespiratory tness. Prog Cardiovasc Dis 63: 341349, 2020.
doi:10.1016/j.pcad.2020.02.001.
724. Williams CJ, Williams MG, Eynon N, Ashton KJ, Little JP, Wisloff U,
Coombes JS.Genes to predict Vo
2max
trainability: a systematic
review. BMC Genomics 18: 831, 2017. doi:10.1186/s12864-017-4192-6.
725. Díaz-Ramírez JD, Álvarez-Herms J, Casta~
neda-Babarro A, Larruskain
J, Ramírez de la Piscina XR, Borisov OV, Semenova EA, Kostryukova
ES, Kulemin NA, Andryushchenko ON, Larin AK, Andryushchenko
LB, Generozov EV, Ahmetov II, Odriozola A.The GALNTL6 Gene
rs558129 polymorphism is associated with power performance.
J Strength Cond Res 34: 30313036, 2020. doi:10.1519/JSC.
0000000000003814.
726. Al-KhelaiF, Yousri NA, Diboun I, Semenova EA, Kostryukova ES,
Kulemin NA, Borisov OV, Andryushchenko LB, Larin AK, Generozov
EV, Miyamoto-Mikami E, Murakami H, Zempo H, Miyachi M,
Takaragawa M, Kumagai H, Naito H, Fuku N, Abraham D, Hingorani
A, Donati F, Botr
e F, Georgakopoulos C, Suhre K, Ahmetov II,
Albagha O, Elrayess MA.Genome-wide association study reveals a
novel association between MYBPC3 gene polymorphism, endur-
ance athlete status, aerobic capacity and steroid metabolism. Front
Genet 11: 595, 2020. doi:10.3389/fgene.2020.00595.
727. Ahmetov I, Kulemin N, Popov D, Naumov V, Akimov E, Bravy Y,
Egorova E, Galeeva A, Generozov E, Kostryukova E, Larin A,
Mustana L, Ospanova E, Pavlenko A, Starnes L, _
Zmijewski P,
Alexeev D, Vinogradova O, Govorun V.Genome-wide association
study identies three novel genetic markers associated with elite
endurance performance. Biol Sport 32: 39, 2015. doi:10.5604/
20831862.1124568.
728. Malczewska-Lenczowska J, Orysiak J, Majorczyk E, Sitkowski D,
Starczewski M, _
Zmijewski P.HIF-1aand NFIA-AS2 polymorphisms
as potential determinants of total hemoglobin mass in endurance
athletes. JStrengthCondRes36: 15961604, 2022. doi:10.1519/
JSC.0000000000003686.
729. Semenova EA, Miyamoto-Mikami E, Akimov EB, Al-KhelaiF,
Murakami H, Zempo H, Kostryukova ES, Kulemin NA, Larin AK,
Borisov OV, Miyachi M, Popov DV, Boulygina EA, Takaragawa M,
Kumagai H, Naito H, Pushkarev VP, Dyatlov DA, Lekontsev EV,
FURRER ET AL.
1782 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
Pushkareva YE, Andryushchenko LB, Elrayess MA, Generozov EV,
Fuku N, Ahmetov II.The association of HFE gene H63D polymor-
phism with endurance athlete status and aerobic capacity: novel
ndings and a meta-analysis. Eur J Appl Physiol 120: 665673,
2020. doi:10.1007/s00421-020-04306-8.
730. Webborn N, Williams A, McNamee M, Bouchard C, Pitsiladis Y,
Ahmetov I, Ashley E, Byrne N, Camporesi S, Collins M, Dijkstra P,
Eynon N, Fuku N, Garton FC, Hoppe N, Holm S, Kaye J, Klissouras
V, Lucia A, Maase K, Moran C, North KN, Pigozzi F, Wang G.Direct-
to-consumer genetic testing for predicting sports performance and
talent identication: consensus statement. Br J Sports Med 49:
14861491, 2015. doi:10.1136/bjsports-2015-095343.
731. G
ullich A, Macnamara BN, Hambrick DZ.What makes a champion?
Early multidisciplinary practice, not early specialization, predicts
world-class performance. Perspect Psychol Sci 17: 629, 2022.
doi:10.1177/1745691620974772.
732. Pickering C, Kiely J.Can genetic testing predict talent? A case
study of 5 elite athletes. Int J Sports Physiol Perform 16: 429434,
2021. doi:10.1123/ijspp.2019-0543.
733. Kujala UM, Leskinen T, Rottensteiner M, Aaltonen S, Ala-Korpela M,
Waller K, Kaprio J.Physical activity and health: ndings from Finnish
monozygotic twin pairs discordant for physical activity. Scand J
Med Sci Sports 32: 13161323, 2022. doi:10.1111/sms.14205.
734. Wang G, Padmanabhan S, Wolfarth B, Fuku N, Lucia A, Ahmetov II,
Cieszczyk P, Collins M, Eynon N, Klissouras V, Williams A, Pitsiladis
Y.Genomics of elite sporting performance: what little we know and
necessary advances. Adv Genet 84: 123149, 2013. doi:10.1016/
B978-0-12-407703-4.00004-9.
735. Sanchis-Gomar F, Pareja-Galeano H, Rodriguez-Marroyo JA, de
Koning JJ, Lucia A, Foster C.Olympic genes on the podium? Int J
Sports Physiol Perform 11: 973974, 2016. doi:10.1123/ijspp.2016-
0421.
736. Urtamo A, Jyv
akorpi SK, Strandberg TE.Denitions of successful
ageing: a brief review of a multidimensional concept. Acta Biomed
90: 359363, 2019. doi:10.23750/abm.v90i2.8376.
737. Seals DR, Justice JN, LaRocca TJ.Physiological geroscience: tar-
geting function to increase healthspan and achieve optimal longev-
ity. J Physiol 594: 20012024, 2016. doi:10.1113/jphysiol.2014.
282665.
738. Lazarus NR, Harridge SD.Declining performance of master ath-
letes: silhouettes of the trajectory of healthy human ageing? J
Physiol 595: 29412948, 2017. doi:10.1113/JP272443.
739. Gries KJ, Trappe SW.The aging athlete: paradigm of healthy aging.
Int J Sports Med 43: 661678, 2022. doi:10.1055/a-1761-8481.
740. Gries KJ, Raue U, Perkins RK, Lavin KM, Overstreet BS, DAcquisto
LJ, Graham B, Finch WH, Kaminsky LA, Trappe TA, Trappe S.
Cardiovascular and skeletal muscle health with lifelong exercise. J
Appl Physiol (1985) 125: 16361645, 2018. doi:10.1152/japplphysiol.
00174.2018.
741. Baker AB, Tang YQ.Aging performance for masters records in ath-
letics, swimming, rowing, cycling, triathlon, and weightlifting. Exp
Aging Res 36: 453477, 2010. doi:10.1080/0361073X.2010.507433.
742. Ganse B, Degens H.Current insights in the age-related decline in
sports performance of the older athlete. Int J Sports Med 42: 879
888, 2021. doi:10.1055/a-1480-7730.
743. GanseB,DreyM,HildebrandF,KnobeM,DegensH.Performance
declines are accelerated in the oldest-old track and eld athletes 80
to 94 years of age. Rejuvenation Res 24: 2027, 2021. doi:10.1089/
rej.2020.2337.
744. Ganse B, Ganse U, Dahl J, Degens H.Linear decrease in athletic
performance during the human life span. Front Physiol 9: 1100,
2018. doi:10.3389/fphys.2018.01100.
745. Ganse B, Kleerekoper A, Knobe M, Hildebrand F, Degens H.
Longitudinal trends in master track and eld performance through-
out the aging process: 83,209 results from Sweden in 16 athletics
disciplines. Geroscience 42: 16091620, 2020. doi:10.1007/s11357-
020-00275-0.
746. Wiswell RA, Hawkins SA, Jaque SV, Hyslop D, Constantino N,
Tarpenning K, Marcell T, Schroeder ET.Relationship between phys-
iological loss, performance decrement, and age in master athletes.
J Gerontol A Biol Sci Med Sci 56: M618M626, 2001. doi:10.1093/
gerona/56.10.m618.
747. Gava P, Ravara B.Master World Records show minor gender differ-
ences of performance decline with aging. Eur J Transl Myol 29:
8327, 2019. doi:10.4081/ejtm.2019.8327.
748. Arampatzis A, Degens H, Baltzopoulos V, Rittweger J.Why do older
sprinters reach the nish line later? Exerc Sport Sci Rev 39: 1822,
2011. doi:10.1097/JES.0b013e318201efe0.
749. Janssen I, Heymseld SB, Wang ZM, Ross R.Skeletal muscle mass
and distribution in 468 men and women aged 18-88 yr. JAppl
Physiol (1985) 89: 8188, 2000. doi:10.1152/jappl.2000.89.1.81.
750. Grosicki GJ, Gries KJ, Minchev K, Raue U, Chambers TL, Begue G,
Finch H, Graham B, Trappe TA, Trappe S.Single muscle bre con-
tractile characteristics with lifelong endurance exercise. J Physiol
599: 35493565, 2021. doi:10.1113/JP281666.
751. Bagley L, McPhee JS, Ganse B, M
uller K, Korhonen MT, Rittweger J,
Degens H.Similar relative decline in aerobic and anaerobic power
with age in endurance and power master athletes of both sexes.
Scand J Med Sci Sports 29: 791799, 2019. doi:10.1111/sms.13404.
752. Cristea A, Korhonen MT, H
akkinen K, Mero A, Al
en M, Sipil
aS,
Viitasalo JT, Koljonen MJ, Suominen H, Larsson L.Effects of com-
bined strength and sprint training on regulation of muscle contrac-
tion at the whole-muscle and single-bre levels in elite master
sprinters. Acta Physiol (Oxf) 193: 275289, 2008. doi:10.1111/j.1748-
1716.2008.01843.x.
753. Grosicki GJ, Standley RA, Murach KA, Raue U, Minchev K, Coen
PM, Newman AB, Cummings S, Harris T, Kritchevsky S, Goodpaster
BH, Trappe S; Health ABC Study. Improved single muscle ber
quality in the oldest-old. J Appl Physiol (1985) 121: 878884, 2016.
doi:10.1152/japplphysiol.00479.2016.
754. Fujimoto N, Hastings JL, Bhella PS, Shibata S, Gandhi NK, Carrick-
Ranson G, Palmer D, Levine BD.Effect of ageing on left ventricular
compliance and distensibility in healthy sedentary humans. J
Physiol 590: 18711880, 2012. doi:10.1113/jphysiol.2011.218271.
755. Beaumont A, Campbell A, Grace F, Sculthorpe N.Cardiac
response to exercise in normal ageing: what can we learn from mas-
ters athletes? Curr Cardiol Rev 14: 245253, 2018. doi:10.2174/
1573403X14666180810155513.
756. Churchill TW, Groezinger E, Kim JH, Loomer G, Guseh JS, Wasfy
MM, Isselbacher EM, Lewis GD, Weiner RB, Schmied C, Baggish AL.
Association of ascending aortic dilatation and long-term endurance
exercise among older masters-level athletes. JAMA Cardiol 5:
522531, 2020. doi:10.1001/jamacardio.2020.0054.
757. Myrstad M, Løchen ML, Graff-Iversen S, Gulsvik AK, Thelle DS,
Stigum H, Ranhoff AH.Increased risk of atrial brillation among
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1783
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
elderly Norwegian men with a history of long-term endurance sport
practice. Scand J Med Sci Sports 24: e238e244, 2014. doi:10.1111/
sms.12150.
758. Newman W, Parry-Williams G, Wiles J, Edwards J, Hulbert S,
Kipourou K, Papadakis M, Sharma R, ODriscoll J.Risk of atrial -
brillation in athletes: a systematic review and meta-analysis. Br
JSportsMed55: 12331238, 2021. doi:10.1136/bjsports-2021-
103994.
759. Momma H, Kawakami R, Honda T, Sawada SS.Muscle-strengthen-
ing activities are associated with lower risk and mortality in major
non-communicable diseases: a systematic review and meta-analy-
sis of cohort studies. Br J Sports Med 56: 755763, 2022.
doi:10.1136/bjsports-2021-105061.
760. Clarke PM, Walter SJ, Hayen A, Mallon WJ, Heijmans J, Studdert
DM.Survival of the ttest: retrospective cohort study of the longev-
ity of Olympic medallists in the modern era. Br J Sports Med 49:
898902, 2015. doi:10.1136/bjsports-2015-e8308rep.
761. Runacres A, Mackintosh KA, McNarry MA.Health consequences of
an elite sporting career: long-term detriment or long-term gain? A
meta-analysis of 165,000 former athletes. Sports Med 51: 289301,
2021. doi:10.1007/s40279-020-01379-5.
762. Lemez S, Baker J.Do elite athletes live longer? A systematic review
of mortality and longevity in elite athletes. Sports Med Open 1: 16,
2015. doi:10.1186/s40798-015-0024-x.
763. Garatachea N, Santos-Lozano A, Sanchis-Gomar F, Fiuza-Luces C,
Pareja-Galeano H, Emanuele E, Lucia A.Elite athletes live longer
than the general population: a meta-analysis. Mayo Clin Proc 89:
11951200, 2014. doi:10.1016/j.mayocp.2014.06.004.
764. Antero J, Tanaka H, De Larochelambert Q, Pohar-Perme M,
Toussaint JF.Female and male US Olympic athletes live 5 years
longer than their general population counterparts: a study of 8124
former US Olympians. Br J Sports Med 55: 206212, 2021.
doi:10.1136/bjsports-2019-101696.
765. Antero-Jacquemin J, Pohar-Perme M, Rey G, Toussaint JF,
Latouche A.The heart of the matter: years-saved from cardiovascu-
lar and cancer deaths in an elite athlete cohort with over a century
of follow-up. Eur J Epidemiol 33: 531543, 2018. doi:10.1007/
s10654-018-0401-0.
766. Lin Y, Gajewski A, Pozna
nska A.Examining mortality risk and rate
of ageing among Polish Olympic athletes: a survival follow-up from
1924 to 2012. BMJ Open 6: e010965, 2016. doi:10.1136/bmjopen-
2015-010965.
767. Marijon E, Tafet M, Antero-Jacquemin J, El Helou N, Berthelot G,
Celermajer DS, Bougouin W, Combes N, Hermine O, Empana JP,
Rey G, Toussaint JF, Jouven X.Mortality of French participants in
theTourdeFrance(1947-2012).Eur Heart J 34: 31453150, 2013.
doi:10.1093/eurheartj/eht347.
768. Kontro TK, Sarna S, Kaprio J, Kujala UM.Mortality and health-
related habits in 900 Finnish former elite athletes and their broth-
ers. Br J Sports Med 52: 8995, 2018. doi:10.1136/bjsports-2017-
098206.
769. Kettunen JA, Kujala UM, Kaprio J, B
ackmand H, Peltonen M,
Eriksson JG, Sarna S.All-cause and disease-specic mortality
among male, former elite athletes: an average 50-year follow-up.
Br J Sports Med 49: 893897, 2015. doi:10.1136/bjsports-2013-
093347.
770. Tang SL.Wearable sensors for sports performance. In: Textiles for
Sportswear, edited by Shishoo R. Cambridge, UK: Woodhead
Publishing, 2015, p. 169196.
771. Passos J, Lopes SI, Clemente FM, Moreira PM, Rico-Gonzaález M,
Bezerra P, Rodrigues LP.Wearables and internet of things (IoT)
technologies for tness assessment: a systematic review. Sensors
(Basel) 21: 5418, 2021. doi:10.3390/s21165418.
772. Cheng Y, Wang K, Xu H, Li T, Jin Q, Cui D.Recent developments in
sensors for wearable device applications. Anal Bioanal Chem 413:
60376057, 2021. doi:10.1007/s00216-021-03602-2.
773. McLaren SJ, Macpherson TW, Coutts AJ, Hurst C, Spears IR,
Weston M.The relationships between internal and external meas-
ures of training load and intensity in team sports: a meta-analysis.
Sports Med 48: 641658, 2018. doi:10.1007/s40279-017-0830-z.
774. Jiang D, Shi G.Research on data security and privacy protection of
wearable equipment in healthcare. J Healthc Eng 2021: 6656204,
2021. doi:10.1155/2021/6656204.
775. Heaney LM, Deighton K, Suzuki T.Non-targeted metabolomics in
sport and exercise science. JSportsSci37: 959967, 2019.
doi:10.1080/02640414.2017.1305122.
776. Khoramipour K, Sandbakk Ø, Keshteli AH, Gaeini AA, Wishart DS,
Chamari K.Metabolomics in exercise and sports: a systematic
review. Sports Med 52: 547583, 2022. doi:10.1007/s40279-021-
01582-y.
777. Morville T, Sahl RE, Moritz T, Helge JW, Clemmensen C.Plasma
metabolome proling of resistance exercise and endurance exer-
cise in humans. Cell Rep 33: 108554, 2020. doi:10.1016/j.
celrep.2020.108554.
778. Deshmukh AS, Steenberg DE, Hostrup M, Birk JB, Larsen JK,
Santos A, Kjøbsted R, Hingst JR, Sch
eele CC, Murgia M, Kiens B,
Richter EA, Mann M, Wojtaszewski JF.Deep muscle-proteomic
analysis of freeze-dried human muscle biopsies reveals ber type-
specic adaptations to exercise training. Nat Commun 12: 304,
2021. doi:10.1038/s41467-020-20556-8.
779. Gomes C, Almeida JA, Franco OL, Petriz B.Omics and the molecular
exercise physiology. Adv Clin Chem 96: 5584, 2020. doi:10.1016/
bs.acc.2019.11.003.
780. Hoffman NJ.Omics and exercise: global approaches for mapping
exercise biological networks. Cold Spring Harb Perspect Med 7:
a029884, 2017. doi:10.1101/cshperspect.a029884.
781. Neufer PD, Bamman MM, Muoio DM, Bouchard C, Cooper DM,
Goodpaster BH, Booth FW, Kohrt WM, Gerszten RE, Mattson MP,
Hepple RT, Kraus WE, Reid MB, Bodine SC, Jakicic JM, Fleg JL,
Williams JP, Joseph L, Evans M, Maruvada P, Rodgers M, Roary M,
Boyce AT, Drugan JK, Koenig JI, Ingraham RH, Krotoski D, Garcia-
Cazarin M, McGowan JA, Laughlin MR.Understanding the cellular
and molecular mechanisms of physical activity-induced health ben-
ets. Cell Metab 22: 411, 2015. doi:10.1016/j.cmet.2015.05.011.
782. Sanford JA, Nogiec CD, Lindholm ME, Adkins JN, Amar D, Dasari S,
et al.Molecular Transducers of Physical Activity Consortium
(MoTrPAC): mapping the dynamic responses to exercise. Cell 181:
14641474, 2020. doi:10.1016/j.cell.2020.06.004.
783. Bongiovanni T, Pintus R, Dessì A, Noto A, Sardo S, Finco G,
Corsello G, Fanos V.Sportomics: metabolomics applied to sports.
The new revolution? Eur Rev Med Pharmacol Sci 23: 1101111019,
2019. doi:10.26355/eurrev_201912_19807.
784. Resende NM, de Magalhães Neto AM, Bachini F, de Castro LE,
Bassini A, Cameron LC.Metabolic changes during a eld experi-
ment in a world-class windsurng athlete: a trial with multivariate
analyses. OMICS 15: 695704, 2011. doi:10.1089/omi.2011.0010.
FURRER ET AL.
1784 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
785. Contrepois K, Wu S, Moneghetti KJ, Hornburg D, Ahadi S, Tsai MS,
et al.Molecular Choreography of Acute Exercise. Cell 181: 1112
1130.e16, 2020. doi:10.1016/j.cell.2020.04.043.
786. Sch
ussler-Fiorenza Rose SM, Contrepois K, Moneghetti KJ, Zhou W,
Mishra T, Mataraso S, Dagan-Rosenfeld O, Ganz AB, Dunn J,
Hornburg D, Rego S, Perelman D, Ahadi S, Sailani MR, Zhou Y,
Leopold SR, Chen J, Ashland M, Christle JW, Avina M, Limcaoco P,
Ruiz C, Tan M, Butte AJ, Weinstock GM, Slavich GM, Sodergren E,
McLaughlin TL, Haddad F, Snyder MP.A longitudinal big data
approach for precision health. Nat Med 25: 792804, 2019.
doi:10.1038/s41591-019-0414-6.
787. Nemkov T, Cendali F, Stefanoni D, Martinez JL, Hansen KC, San-
Millán I, DAlessandro A.Metabolic signatures of performance in elite
world tour professional cyclists (Preprint). bioRxiv 2022.09.13.507793,
2022. doi:10.1101/2022.09.13.507793.
788. Tanaka M, Wang G, Pitsiladis YP.Advancing sports and exercise
genomics: moving from hypothesis-driven single study approaches
to large multi-omics collaborative science. Physiol Genomics 48:
173174, 2016. doi:10.1152/physiolgenomics.00009.2016.
789. Lindinger MI, Ward SA.A century of exercise physiology: key con-
cepts in....Eur J Appl Physiol 122: 14, 2022. doi:10.1007/s00421-
021-04873-4.
790. Karstoft K, Pedersen BK.Skeletal muscle as a gene regulatory en-
docrine organ. Curr Opin Clin Nutr Metab Care 19: 270275, 2016.
doi:10.1097/MCO.0000000000000283.
791. Whitham M, Febbraio MA.The ever-expanding myokinome: discov-
ery challenges and therapeutic implications. Nat Rev Drug Discov
15: 719729, 2016. doi:10.1038/nrd.2016.153.
792. Cervenka I, Agudelo LZ, Ruas JL.Kynurenines: tryptophansmetab-
olites in exercise, inammation, and mental health. Science 357:
eaaf9794, 2017. doi:10.1126/science.aaf9794.
793. Dos Santos M, Backer S, Saintpierre B, Izac B, Andrieu M,
Letourneur F, Relaix F, Sotiropoulos A, Maire P.Single-nucleus
RNA-seq and FISH identify coordinated transcriptional activity in
mammalian myobers. Nat Commun 11: 5102, 2020. doi:10.1038/
s41467-020-18789-8.
794. Kim M, Franke V, Brandt B, Lowenstein ED, Sch
owel V, Spuler S,
Akalin A, Birchmeier C.Single-nucleus transcriptomics reveals func-
tional compartmentalization in syncytial skeletal muscle cells. Nat
Commun 11: 6375, 2020. doi:10.1038/s41467-020-20064-9.
795. Petrany MJ, Swoboda CO, Sun C, Chetal K, Chen X, Weirauch MT,
Salomonis N, Millay DP.Single-nucleus RNA-seq identies tran-
scriptional heterogeneity in multinucleated skeletal myobers. Nat
Commun 11: 6374, 2020. doi:10.1038/s41467-020-20063-w.
796. Crudele JM, Chamberlain JS.AAV-based gene therapies for the
muscular dystrophies. Hum Mol Genet 28: R102R107, 2019.
doi:10.1093/hmg/ddz128.
797. Mendell JR, Al-Zaidy SA, Rodino-Klapac LR, Goodspeed K, Gray SJ,
Kay CN, Boye SL, Boye SE, George LA, Salabarria S, Corti M, Byrne
BJ, Tremblay JP.Current clinical applications of in vivo gene ther-
apy with AAVs. Mol Ther 29: 464488, 2021. doi:10.1016/j.
ymthe.2020.12.007.
798. Pedersen BK.The physiology of optimizing health with a focus on
exercise as medicine. Annu Rev Physiol 81: 607627, 2019.
doi:10.1146/annurev-physiol-020518-114339.
799. Pedersen BK, Saltin B.Exercise as medicineevidence for pre-
scribing exercise as therapy in 26 different chronic diseases. Scand
J Med Sci Sports 25, Suppl 3: 172, 2015. doi:10.1111/sms.12581.
800. Ganse B, Degens H.Declining track and eld performance trends in
recent years in the Austrian best results 1897-2019. J Musculoskelet
Neuronal Interact 21: 196205, 2021.
801. Nugent FJ, Comyns TM, Warrington GD.Quality versus quantity
debate in swimming: perceptions and training practices of expert
swimming coaches. J Hum Kinet 57: 147158, 2017. doi:10.1515/
hukin-2017-0056.
802. Crowley E, Harrison AJ, Lyons M.The impact of resistance training
on swimming performance: a systematic review. Sports Med 47:
22852307, 2017. doi:10.1007/s40279-017-0730-2.
803. Crowley E, Ng K, Mujika I, Powell C.Speeding up or slowing down?
Analysis of race results in elite-level swimming from 2011-2019 to
predict future Olympic Games performances. Meas Phys Educ
Exerc Sci 26: 130140, 2022. doi:10.1080/1091367X.2021.1952592.
804. Robinson AT, Watso JC, Babcock MC, Joyner MJ, Farquhar WB.
Record-breaking performance in a 70-year-old marathoner. N Engl
JMed380: 14851486, 2019. doi:10.1056/NEJMc1900771.
805. Moore IS, Willy RW.Use of wearables: tracking and retraining in en-
durance runners. Curr Sports Med Rep 18: 437444, 2019.
doi:10.1249/JSR.0000000000000667.
806. Laranjo L, Ding D, Heleno B, Kocaballi B, Quiroz JC, Tong HL,
Chahwan B, Neves AL, Gabarron E, Dao KP, Rodrigues D, Neves
GC, Antunes ML, Coiera E, Bates DW.Do smartphone applications
and activity trackers increase physical activity in adults? Systematic
review, meta-analysis and metaregression. Br J Sports Med 55:
422432, 2021. doi:10.1136/bjsports-2020-102892.
807. Murach KA, McCarthy JJ, Peterson CA, Dungan CM.Making mice
mighty: recent advances in translational models of load-induced
muscle hypertrophy. J Appl Physiol (1985) 129: 516521, 2020.
doi:10.1152/japplphysiol.00319.2020.
808. Poole DC, Copp SW, Colburn TD, Craig JC, Allen DL, Sturek M,
OLeary DS, Zucker IH, Musch TI.Guidelines for animal exercise
and training protocols for cardiovascular studies. Am J Physiol
Heart Circ Physiol 318: H1100H1138, 2020. doi:10.1152/ajpheart.
00697.2019.
809. Ferreira-Pinto MJ, Ruder L, Capelli P, Arber S.Connecting circuits
for supraspinal control of locomotion. Neuron 100: 361374, 2018.
doi:10.1016/j.neuron.2018.09.015.
810. Thornton JS, Patricios J, Kemp J, Engebretsen L, Drezner J.Robin
Hood in SEM? What can we take from elite sport to give back to
wider public health? Br J Sports Med 55: 949950, 2021.
doi:10.1136/bjsports-2021-104266.
811. Millet GP, Seiler S, Millet GY.Opportunities and obstacles of trans-
lating elite sport research to public health. Br J Sports Med 56: 64
65, 2022. doi:10.1136/bjsports-2021-104374.
812. Billat LV.Interval training for performance: a scientic and empirical
practice. Special recommendations for middle- and long-distance
running. Part I: aerobic interval training. Sports Med 31: 1331, 2001.
doi:10.2165/00007256-200131010-00002.
813. Delezie J, Handschin C.Endocrine crosstalk between skeletal mus-
cle and the brain. Front Neurol 9: 698, 2018. doi:10.3389/fneur.
2018.00698.
814. Nay K, Smiles WJ, Kaiser J, McAloon LM, Loh K, Galic S, Oakhill JS,
Gundlach AL, Scott JW.Molecular mechanisms underlying the ben-
ecial effects of exercise on brain function and neurological disor-
ders. Int J Mol Sci 22: 4052, 2021. doi:10.3390/ijms22084052.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1785
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
815. Heuberger J, Cohen AF.Review of WADA prohibited substances:
limited evidence for performance-enhancing effects. Sports Med
49: 525539, 2019. doi:10.1007/s40279-018-1014-1.
816. Guerrieri D, Moon HY, van Praag H.Exercise in a pill: the latest on
exercise-mimetics. Brain Plast 2: 153169, 2017. doi:10.3233/BPL-
160043.
817. Vara-Ciruelos D, Russell FM, Hardie DG.The strange case of
AMPK and cancer: Dr Jekyll or Mr Hyde? Open Biol 9: 190099,
2019. doi:10.1098/rsob.190099.
818. Wackerhage H, Schoenfeld BJ.Evidence-informed training plans
and exercise prescriptions for performance, tness and health.
Sports Med 51: 18051813, 2021. doi:10.1007/s40279-021-01495-w.
819. Grindem H, Myklebust G.Be a champion for your athleteshealth.J
Orthop Sports Phys Ther 50: 173175, 2020. doi:10.2519/jospt.2020.
0605.
820. Tiller NB, Sullivan JP, Ekkekakis P.Baseless claims and pseudo-
science in health and wellness: a call to action for the sports, exer-
cise, and nutrition-science community. Sports Med 53: 15, 2023.
doi:10.1007/s40279-022-01702-2.
821. McCarthy O, Pitt JP, Keay N, Vestergaard ET, Tan AS, Churm R,
Rees DA, Bracken RM.Passing on the exercise baton: what can en-
docrine patients learn from elite athletes? Clin Endocrinol (Oxf) 96:
781792, 2022. doi:10.1111/cen.14683.
822. Millet GP, Chamari K.Look to the starsis there anything that
public health and rehabilitation can learn from elite sports? Front
Sports Act Living 4: 1072154, 2023. doi:10.3389/fspor.2022.
1072154.
823. Lightfoot JT, DE Geus EJ, Booth FW, Bray MS, DEN Hoed M, Kaprio
J, Kelly SA, Pomp D, Saul MC, Thomis MA, Garland T Jr, Bouchard
C.Biological/genetic regulation of physical activity level: consensus
from GenBioPAC. Med Sci Sports Exerc 50: 863873, 2018.
doi:10.1249/MSS.0000000000001499.
824. Morgan PJ, Young MD, Smith JJ, Lubans DR.Targeted health
behavior interventions promoting physical activity: a conceptual
model. Exerc Sport Sci Rev 44: 7180, 2016. doi:10.1249/JES.
0000000000000075.
825. Sherwood NE, Jeffery RW.The behavioral determinants of exer-
cise: implications for physical activity interventions. Annu Rev Nutr
20: 2144, 2000. doi:10.1146/annurev.nutr.20.1.21.
826. Tao G, Garrett B, Taverner T, Cordingley E, Sun C.Immersive virtual
reality health games: a narrative review of game design. J
Neuroeng Rehabil 18: 31, 2021. doi:10.1186/s12984-020-00801-3.
827. Berg J, Haugen G, Wang AI, Moholdt T.High-intensity exergaming
for improved cardiorespiratory tness: a randomised, controlled
trial. Eur J Sport Sci 22: 867876, 2022. doi:10.1080/17461391.
2021.1921852.
828. Schultz MG, Sharman JE.Exercise hypertension. Pulse (Basel) 1:
161176, 2014. doi:10.1159/000360975.
829. W
urzburger L, Wiech P, Rossi VA, Neunh
auserer D, Caselli S,
Schmied CM, Niederseer D.Hypertensive response to exercise in
athletes: unremarkable nding or relevant marker for future cardio-
vascular complications? Int J Hypertens 2022: 8476751, 2022.
doi:10.1155/2022/8476751.
830. Giulianotti R, Thiel A.New horizons in the sociology of sport. Front
Sports Act Living 4: 1060622, 2023. doi:10.3389/fspor.2022.1060622.
831. Schoenfeld B, Fisher J, Grgic J, Haun C, Helms E, Phillips S, Steele
J, Vigotsky A.Resistance training recommendations to maximize
muscle hypertrophy in an athletic population: position stand of the
IUSCA. Int J Strength Cond 1: 1, 2021. doi:10.47206/ijsc.v1i1.81.
832. Elder GC, Bradbury K, Roberts R.Variability of ber type distribu-
tions within human muscles. J Appl Physiol Respir Environ Exerc
Physiol 53: 14731480, 1982. doi:10.1152/jappl.1982.53.6.1473.
833. Wan JJ, Qin Z, Wang PY, Sun Y, Liu X.Muscle fatigue: general
understanding and treatment. Exp Mol Med 49: e384, 2017.
doi:10.1038/emm.2017.194.
834. Loenneke JP.Muscle growth does not contribute to the increases in
strength that occur after resistance training. Med Sci Sports Exerc
53: 20112014, 2021. doi:10.1249/MSS.0000000000002662.
835. Folland JP, Balshaw TG.muscle growth does contribute to the
increases in strength that occur after resistance training. Med
Sci Sports Exerc 53: 20062010, 2021. doi:10.1249/MSS.
0000000000002732.
836. Mendell LM.The size principle: a rule describing the recruitment of
motoneurons. JNeurophysiol93: 30243026, 2005. doi:10.1152/
classicessays.00025.2005.
837. Alkaslasi MR, Piccus ZE, Hareendran S, Silberberg H, Chen L,
Zhang Y, Petros TJ, Le Pichon CE.Single nucleus RNA-sequencing
denes unexpected diversity of cholinergic neuron types in the
adult mouse spinal cord. Nat Commun 12: 2471, 2021. doi:10.1038/
s41467-021-22691-2.
838. Liu JX, H
oglund AS, Karlsson P, Lindblad J, Qaisar R, Aare S,
Bengtsson E, Larsson L.Myonuclear domain size and myosin iso-
form expression in muscle bres from mammals representing a
100,000-fold difference in body size. Exp Physiol 94: 117129,
2009. doi:10.1113/expphysiol.2008.043877.
839. Gundersen K.Muscle memory and a new cellular model for muscle
atrophy and hypertrophy. JExpBiol219: 235242, 2016. doi:10.1242/
jeb.124495.
840. Schwartz LM.Skeletal muscles do not undergo apoptosis during ei-
ther atrophy or programmed cell deathrevisiting the myonuclear
domain hypothesis. Front Physiol 9: 1887, 2018. doi:10.3389/
fpls.2018.01887.
841. Gundersen K, Bruusgaard JC.Nuclear domains during muscle atro-
phy: nuclei lost or paradigm lost? J Physiol 586: 26752681, 2008.
doi:10.1113/jphysiol.2008.154369.
842. Schwartz LM, Gundersen K.Cross Talk opposing view: myonuclei
do not undergo apoptosis during skeletal muscle atrophy. J Physiol
600: 20812084, 2022. doi:10.1113/JP282381.
843. Kirby TJ, Dupont-Versteegden EE.Cross Talk proposal: myonuclei
are lost with ageing and atrophy. J Physiol 600: 20772080, 2022.
doi:10.1113/JP282380.
844. Van der Meer SF, Jaspers RT, Degens H.Is the myonuclear domain
size xed? J Musculoskelet Neuronal Interact 11: 286297, 2011.
845. Murach KA, Englund DA, Dupont-Versteegden EE, McCarthy JJ,
Peterson CA.Myonuclear domain exibility challenges rigid
assumptions on satellite cell contribution to skeletal muscle ber
hypertrophy. Front Physiol 9: 635, 2018. doi:10.3389/fphys.2018.
00635.
846. Aman F, El Khatib E, AlNeaimi A, Mohamed A, Almulla AS, Zaidan A,
Alshafei J, Habbal O, Eldesouki S, Qaisar R.Is the myonuclear domain
ceiling hypothesis dead? Singapore Med J. In Press. doi:10.11622/
smedj.2021103.
FURRER ET AL.
1786 Physiol Rev VOL 103 JULY 2023 www.prv.org
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
847. Azevedo M, Baylies MK.Getting into position: nuclear movement in
muscle cells. Trends Cell Biol 30: 303316, 2020. doi:10.1016/j.
tcb.2020.01.002.
848. Cavanagh JB.The problems of neurons with long axons. Lancet 1:
12841287, 1984. doi:10.1016/s0140-6736(84)92457-7.
849. Wickiewicz TL, Roy RR, Powell PL, Edgerton VR.Muscle architec-
ture of the human lower limb. Clin Orthop Relat Res 179: 275283,
1983.
850. BeetsMW, Weaver RG, Brazendale K.Daring to share requires inten-
tional and collective commitment to civil discourse. Int J Behav Nutr
Phys Act 17: 46, 2020. doi:10.1186/s12966-020-00950-7.
851. Ledgerwood A.What do we want our scientic discourse to look
like? Observer 30: 1819, 2017.
852. Sun Q, Wojcieszak M, Davidson S.Over-time trends in incivility
on social media: evidence from political, non-political, and mixed
sub-reddits over eleven years. Front Polit Sci 3: e741605, 2021.
THE MOLECULAR ATHLETE: FROM MOLECULES TO MEDALS
Physiol Rev VOL 103 JULY 2023 www.prv.org 1787
Downloaded from journals.physiology.org/journal/physrev at University of Basel (131.152.009.037) on May 26, 2023.
... The blood lactate threshold increases by approximately 17 % after 40 weeks of training [11]. Individuals gradually increase the training workload by manipulating the intensity, frequency, or duration of training sessions to maximise improvements in V O 2peak and the blood lactate inflection exercise point [12]. The gradual increase in training workload is termed "progressive overload" [12]. ...
... Individuals gradually increase the training workload by manipulating the intensity, frequency, or duration of training sessions to maximise improvements in V O 2peak and the blood lactate inflection exercise point [12]. The gradual increase in training workload is termed "progressive overload" [12]. The theory for progressive overload is modified from the general adaption syndrome theory proposed by Seyle [13]. ...
... The progressive overload theory is based on the theory that adaptation ("supra-compensation") occurs in response to stress to ultimately improve the individual's physiological state and capacity for exercise. Exercise disturbs normal homeostatic regulation as cell ATP concentration, glycogen concentration and pH fall, metabolites accumulate and electrolyte concentration, hormone and immune regulation is altered as is blood and oxygen supply [8,12]. These physiological disturbances are more profound with higher exercise intensity and trigger acute genetic and molecular responses to maintain or re-establish homeostasis [8,12]. ...
... • Furrer (Furrer et al., 2023) cautions that we still do not know what training really works best; as there is little longitudinal research and periodisation is scientifically on weak foundations. • For all-year or multi-year training plans there is limited evidence whether the effects of block periodised training increase, maintain, or diminish over the long-term period of time (Kataoka et al., 2021). ...
... Periodisation of training for an elite athlete(Furrer, 2023) ...
Thesis
Full-text available
Amateur marathon runners desire to excel at their chosen sport and to use the correct and latest research on how to optimise training and competition outcomes. Yet they do not have access to professional team of sport scientists, nutritionists, psychologists, and a well-equipped sports lab. This paper intends to review from the perspective of a self-coached 50year old sub-3 marathon runner for the marathon world championships in Sydney what is the latest research, what tools and technologies are available and how can they be integrated into the training of such amateur athletes. We will construct an Annual Training Plan; our starting point is the current level within a multi-year plan, the race ambitions for current year, and overall longer-term athletic goals. With an amateur runner, five years of experience, and a desire to diversify into middle-distance triathlon for current athletic year we will use a traditional training plan with a single peak for Sydney. We now need to assess the athletes fitness, based on VO2max , LT, and RE. Based on this, the desired race outcomes, and the training phase we construct the weekly running workouts. Generally speaking VO 2max is best supported by HIIT, LT by long-runs, and RE by running volume and strength workouts. The initial stages focus on volume which is gradually replaced by intensity. To help recovery within the week we need to vary training intensity and include lower-intensity weeks. To reduce injury risk from high running mileage we will focus only on four high quality run workouts and enhance the overall aerobic system with cross-training (cycling and swimming) and strength training. Training intensity needs to be distributed in a polarised way with 80% of volume in moderate aerobic zones and upto 20% in severe zones. Strength training is 3 session in the general stages, becoming 2 in competitive stage and completely removed in final pre-race weeks. To “Dose & Response” running intensity we will use critical power as measured by Stryd and benchmark against gold-standard laboratory tests. Other measures we will track with TrainingPeaks and WKO5 are key internal and external load for stress management (CTL, RHR, HRV), recovery management (sleep duration, time awake, and sleep quality), injury prevention (weekly running mileage and perceived injury for injury prevention), body composition (caloric expenditure, BMR, BM, and skinfold measurement), and polarised weekly running volume (hours of run training by intensity domain). Nutrition needs to be aligned to daily workloads mostly by varying levels of CHO, while guaranteeing a steady and well-distributed level of protein, mostly in the form of EAA. Nutritional needs are aligned with phases in the training plan, most notably in later precompetition stages where glycogen stores need to be topped-up while keeping body mass as low as possible. This is also supported by 1-2 LSD in fasted state to help use fat as substrate. In addition to physiologic adaptations, training also needs to hone psychological skills. Mental fatigue can be detrimental to competition as central and peripheral muscle fatigue. We will train the psychology by focusing on setting goals, dissociation from fatigue, association with the flow of the exercise, attentional focus which we will also train through yoga, visualisation of the event, positive self-talk, and flow & prayer. With 12 time zone difference and 24-hour long flight we must minimise travel fatigue and jet lag with melatonin, nutrition, recovery, easy workouts, and pre-taper psychology. Moreover we need gut training to assume 90 g/CHO/hour with 40mg caffeine. If all this is executed well we anticipate a sub-3, fastest Hungarian running outcome for Sydney 2024 world championships.
... Numerous positive health and physical performance adaptations have been reported when the stressful stimulus of resistance exercise is administered with appropriate time intervals and optimal recovery time in different phenotypes [54][55][56][57][58][59]. This remarkable plasticity is responsible for adaptations in physiological and pathophysiological conditions [60][61][62][63]. This ability of muscles to adapt to external stimuli relies on different signaling pathways and secondary messengers, which are responsible for maintaining muscle status. ...
Article
Full-text available
The ketogenic diet (KD) is a nutritional strategy characterized by a reduced intake of carbohydrates (between 30 and 45 g per day or ≈5% of one's total calories from this macronutrient). The regimen induces physiological ketosis in which serum levels of ketone bodies increase from 0.5 to 3.0 mM, becoming an essential contributor to energy production. The popularity of using the KD to lose weight and its application in specific physio-pathological conditions, such as epilepsy, lipedema, and polycystic ovary syndrome, which is maintained over extended periods, gave us the impulse to write this brief review. In these types of physio-pathological conditions, subjects can achieve favorable training outcomes even if adhering to a KD. Therefore, performing resistance training under the KD to enhance muscle status and quality of life could be possible. It is important to note that, while some statements here suggest potential future directions, they are hypotheses that require experimental validation, even if they are supported by the independent benefits reported from the KD and resistance training and represent a promising area for future research.
... Studies employing the herbal remedy "Body Fu Kang" have demonstrated that this remedy can considerably improve the skeletal muscles' energy supply and successfully reduce the structural alterations in muscle fibres brought on by exercise stress [32]. Furrer, Hawley [33] employed pressure pain, the extent of healing following exercise, and the occurrence of repeat injuries as indications of efficacy when treating skeletal muscle strain in quadriceps and hamstring injuries using herbal Huangqi and Danshen injections. According to the findings, the injection's topical application may help skeletal muscle injuries heal to some extent and reduce the risk of reinjury after treatment. ...
Article
Full-text available
Different parts of many plants, including seeds, bark, leaves, roots, fruit, stems, or flowers with known or suspected therapeutic properties are used to make herbal medications. In the past ten years, the number of athletes using herbal supplements has increased dramatically. Herbal remedies are becoming more and more popular among athletes and non-athletes as a way of improving their endurance and strength. Several diseases and impairments related to body stress are managed using herbal adaptogens; these adaptogens are also used to enhance focus, boost endurance during fatigue moments, improve physical strength/stamina, enhance energy levels, restore stress-affected cognitive function, improve sexual dysfunction, and maintain the level of cortisol. This study employed a research approach that requires the use of terms like “Herbal adaptogens, ashwagandha, endurance, athletes, turmeric, muscle strength” during a preliminary search of some of the popular databases such as Google, PubMed, Embase, ScienceDirect, OVID Medline, Google Scholar, and Web of Science. The leading herbal adaptogens on the global market (such as ashwagandha , Rhodiola roseas , astragalus, holy basil, cordyceps, and turmeric) were examined in this article based on their source. Also covered in this work are the potential negative effects of these adaptogens and how they can help athletes perform better by increasing their muscle mass, recovery, and endurance.
... This improvement is attributed to elevated aerobic enzyme concentration and increased mitochondrial function [37], size, number, and surface area [38], as they contribute to improved oxygen extraction by the working muscles. Moreover, high aerobic fitness is associated with increased muscle blood flow, capillarization of muscle tissue, blood, and total hemoglobin volume, which improves oxygen delivery as well as lactate metabolism and transport [39]. Consequently, more energy is supplied through the aerobic and phosphagen systems with decreased contribution of anaerobic glycolysis. ...
Article
Full-text available
Respiratory muscle training plays a significant role in reducing blood lactate concentration (bLa) and attenuating negative physiological stress reactions. Therefore, we investigated if voluntary isocapnic hyperpnoea (VIH) performed after a maximum anaerobic effort influences bLa and perceived fatigue level in well-trained speedskaters. 39 elite short-track speedskaters participated in a trial with two parallel groups: experimental and control. All the participants performed the Wingate Anaerobic Test (WAnT). The experimental group performed a VIH-based recovery protocol 20 min after exercise, the control group used passive recovery only. Blood samples were taken 3 and 30 min after the WAnT to measure bLa. Fatigue was self-appraised on a 0–10 perceived rating-of-fatigue (ROF) scale 3 and 30 min after the WAnT. Noteworthy, but not statistically significant changes between the experimental and control groups were observed for changes in bLa (p = 0.101). However, statistically significant changes between the groups were found for ROF (p = 0.003, ηp² = 0.211, ω² = 0.106). Moreover, statistically significant interactions between post-exercise bLa clearance and VO2max (p = 0.028) and inspiratory muscle strength (p = 0.040) were observed. Our findings provided preliminary insight that VIH may be an efficient recovery protocol after anaerobic exercise performed by elite athletes. The association between VO2max and post-exercise bLa clearance indicates the vital role of aerobic fitness in repeated-efforts ability in short-track speedskaters. The study was registered at ClinicalTrials.gov as NCT05994092 on 15th August 2023.
... Therefore, practicing vigorous physical activity is more efficient in improving muscle strength compared to moderate activities (Smith et al., 2019). Several aspects may be related to the absence of this association between vigorous physical activity and HGS in boys, such as the specificity of the physical activities performed and may not be appropriate for improving strength, insufficient volume or low sensitivity to the stimuli provided (Furrer et al., 2023;Landry & Driscoll, 2012). However, in relation to volume, boys spend more time in vigorous activities than girls (p < 0.05), thus, the lack of association may be due to the benefits provided having already been incorporated. ...
... Disruptions to this clock machinery can lead to detrimental effects on muscle function and overall metabolic homeostasis. Exercise, a potent modulator of the skeletal muscle clock, has emerged as a promising candidate for entraining the circadian clock, thereby contributing to the prevention and treatment of individuals with metabolic diseases (Furrer et al., 2023;Gabriel & Zierath, 2019;Mansingh & Handschin, 2022;Xin et al., 2023). Recent works have extensively characterized the coordinated expression of the core clock factors BMAL1 and CLOCK within the muscle and identified their interaction with the muscle-specific transcription factor MYOD1 as a facilitator of the circadian and metabolic programmes that support skeletal muscle physiology (reviewed extensively by Martin et al., 2023b). ...
Article
Full-text available
Circadian rhythms, governed by the dominant central clock, in addition to various peripheral clocks, regulate almost all biological processes, including sleep–wake cycles, hormone secretion and metabolism. In certain contexts, the regulation and function of the peripheral oscillations can be decoupled from the central clock. However, the specific mechanisms underlying muscle‐intrinsic clock‐dependent modulation of muscle function and metabolism remain unclear. We investigated the outcome of perturbations of the primary and secondary feedback loops of the molecular clock in skeletal muscle by specific gene ablation of Period circadian regulator 2 (Per2) and RAR‐related orphan receptor alpha (Rorα), respectively. In both models, a dampening of core clock gene oscillation was observed, while the phase was preserved. Moreover, both loops seem to be involved in the homeostasis of amine groups. Highly divergent outcomes were seen for overall muscle gene expression, primarily affecting circadian rhythmicity in the PER2 knockouts and non‐oscillating genes in the RORα knockouts, leading to distinct outcomes in terms of metabolome and phenotype. These results highlight the entanglement of the molecular clock and muscle plasticity and allude to specific functions of different clock components, i.e. the primary and secondary feedback loops, in this context. The reciprocal interaction between muscle contractility and circadian clocks might therefore be instrumental to determining a finely tuned adaptation of muscle tissue to perturbations in health and disease. image Key points Specific perturbations of the primary and secondary feedback loop of the molecular clock result in specific outcomes on muscle metabolism and function. Ablation of Per2 (primary loop) or Rorα (secondary loop) blunts the amplitude of core clock genes, in absence of a shift in phase. Perturbation of the primary feedback loop by deletion of PER2 primarily affects muscle gene oscillation. Knockout of RORα and the ensuing modulation of the secondary loop results in the aberrant expression of a large number of non‐clock genes and proteins. The deletion of PER2 and RORα affects muscle metabolism and contractile function in a circadian manner, highlighting the central role of the molecular clock in modulating muscle plasticity.
Chapter
A significant fundamental change in healthcare can be seen by the integration of artificial intelligence (AI) and machine learning (ML) with medical imaging (MI). This integration holds the potential to improve patient care standards, change clinical procedures, and improve diagnostic accuracy. This chapter presents the possible impact, addressed challenges, opportunities, and impacts on medical diagnosis of recent advancements and growing advances in the application of AI and ML in the field of MI. The rapid development of AI and ML technologies have pushed MI into a new phase of data-driven healthcare and personalized medicine. Furthermore, the applications of AI and ML in MI are a variety of including automated picture analysis and disease identification to prediction, revolutionizing pathology and radiology, and healthcare management. It also suggests a future in which these innovations will work together to improve patient care, improve diagnostic accuracy, and advance healthcare delivery.
Article
Full-text available
Background and Objective Metabolomics studies of recreational and elite athletes have been so far limited to venipuncture-dependent blood sample collection in the setting of controlled training and medical facilities. However, limited to no information is currently available to determine if findings in laboratory settings are translatable to a real-world scenario in elite competitions. The goal of this study was to define molecular signatures of exertion under controlled exercise conditions and use these signatures as a framework for assessing cycling performance in a World Tour competition. Methods To characterize molecular profiles of exertion in elite athletes during cycling, we performed metabolomics analyses on blood isolated from 28 international-level, elite, World Tour professional male athletes from a Union Cycliste Internationale World Team taken before and after a graded exercise test to volitional exhaustion and before and after a long aerobic training session. Moreover, established signatures were then used to characterize the metabolic physiology of five of these cyclists who were selected to represent the same Union Cycliste Internationale World Team during a seven-stage elite World Tour race. Results Using dried blood spot collection to circumvent logistical hurdles associated with field sampling, these studies defined metabolite signatures and fold change ranges of anaerobic or aerobic exertion in elite cyclists, respectively. Blood profiles of lactate, carboxylic acids, fatty acids, and acylcarnitines differed between exercise modes. The graded exercise test elicited significant two- to three-fold accumulations in lactate and succinate, in addition to significant elevations in free fatty acids and acylcarnitines. Conversely, the long aerobic training session elicited a larger magnitude of increase in fatty acids and acylcarnitines without appreciable increases in lactate or succinate. Comparable signatures were revealed after sprinting and climbing stages, respectively, in a World Tour race. In addition, signatures of elevated fatty acid oxidation capacity correlated with competitive performance. Conclusions Collectively, these studies provide a unique view of alterations in the blood metabolome of elite athletes during competition and at the peak of their performance capabilities. Furthermore, they demonstrate the utility of dried blood sampling for omics analysis, thereby enabling molecular monitoring of athletic performance in the field during training and competition.
Article
Full-text available
The relevance of a sociological view on the problems of society has never been as important as it is today. To quote the editors of the journal Nature in their editorial, Time for the Social Sciences, from 2015: if you want science to deliver for society, you need to support a capacity to understand that society. In other words, the technological and scientific disciplines cannot simply transfer their findings into everyday life without knowing how society works. But this realisation does not seem to have caught on everywhere. The sociology of sport is entering a critical period that will shape its development and potential transformation over the next decade. In this paper, we review key features and trends within the sociology of sport in recent times, and set out potential future challenges and ways forward for the subdiscipline. Accordingly, our discussion spans a wide range of issues concerning the sociology of sport, including theories and approaches, methods, and substantive research topics. We also discuss the potential contributions of the sociology of sport to addressing key societal challenges. To examine these issues, the paper is organized into three main parts. First, we identify three main concentric challenges, or types of peripheral status, that sociologists of sport must confront: as social scientists, as sociologists, and as sociologists of sport, respectively. Second, we consider various strengths within the positions of sociology and the sociology of sport. Third, in some detail, we set out several ways forward for the sociology of sport with respect to positioning within academe, scaling up research, embracing the glocal and cosmopolitan aspects of sociology, enhancing plurality in theory, improving transnational coordination, promoting horizontal collaborations, and building greater public engagement. The paper is underpinned by over 60 years (combined) of work within the sociology of sport, including extensive international research and teaching.
Article
Full-text available
This article lays out the determinants of maximal O2 consumption (VO2max) achieved during high intensity endurance exercise. It is not a traditional topical review but rather an educational essay that intertwines chance observations made during an unrelated research project with a subsequent program of stepwise thought, analysis and experimentation to reveal how O2 is delivered to and used by the mitochondria. The centerpiece is the recognition that O2 is delivered by an inter-dependent system of transport components functioning as a “bucket brigade”, made up of the lungs, heart, blood and circulation, and the muscles themselves, each of which affects O2 transport by similar amounts as they change. There is thus no single “limiting factor” to VO2max. Moreover, each component is shown to quantitatively affect the performance of the others. Mitochondrial respiration is integrated into the O2 transport system analysis to reveal its separate contribution to VO2max, and to show that mitochondrial PO2 at VO2max must be extremely low. Clinical application of the O2 transport systems analysis is described to separate central cardiopulmonary from peripheral tissue contributions to exercise limitation, illustrated by a study of patients with COPD. Finally, a short discussion of why muscles operating maximally must endure an almost anoxic state is offered. The hope is that in sum, both the increased understanding of O2 transport and the scientific approach to achieving that understanding described in the review can serve as a model for solving other complex problems going forward.
Article
Full-text available
Understanding how movement is controlled by the CNS remains a major challenge, with ongoing debate about basic features underlying this control. In current established views, the concepts of motor neuron recruitment order, common synaptic input to motor neurons and muscle synergies are usually addressed separately and therefore seen as independent features of motor control. In this review, we analyse the body of literature in a broader perspective and we identify a unified approach to explain apparently divergent observations at different scales of motor control. Specifically, we propose a new conceptual framework of the neural control of movement, which merges the concept of common input to motor neurons and modular control, together with the constraints imposed by recruitment order. This framework is based on the following assumptions: (1) motor neurons are grouped into functional groups (clusters) based on the common inputs they receive; (2) clusters may significantly differ from the classical definition of motor neuron pools, such that they may span across muscles and/or involve only a portion of a muscle; (3) clusters represent functional modules used by the CNS to reduce the dimensionality of the control; and (4) selective volitional control of single motor neurons within a cluster receiving common inputs cannot be achieved. Here, we discuss this framework and its underlying theoretical and experimental evidence. image
Article
Full-text available
Background: In the general population, hypertensive response to exercise (HRE) predicts new-onset resting hypertension or other cardiovascular diseases. Methods: PubMed was searched for English articles published between January 1st 2000 and April 30th 2020. Additional studies were identified via reference lists of included studies. 92 papers were selected for full text analysis, finally 30 studies were included. Results: The results from 5 follow-up studies suggested an association between HRE and the risk of developing hypertension, while 10 studies reported a link with adverse cardiovascular events in the general population. Another study showed an association between HRE and future hypertension in athletes after a follow-up of 7 years. HRE in athletes was associated with left ventricular hypertrophy in three studies. Two other studies showed a link between HRE and focal myocardial fibrosis in triathletes and myocardial injury, respectively. One study found lower Apoliprotein-1 serum levels in athletes with HRE leading to a higher risk for cardiovascular disease. Only in one study no association with cardiovascular dysfunction in athletes with HRE was found. Conclusions: Based on current evidence, HRE is not a normal finding in athletes. If detected, it should be interpreted as a risk factor for future cardiovascular complications. Future research should address the adequate follow-up and management of athletes with HRE.
Article
Full-text available
Voluntary movement requires communication from cortex to the spinal cord, where a dedicated pool of motor units (MUs) activates each muscle. The canonical description of MU function rests upon two foundational tenets. First, cortex cannot control MUs independently but supplies each pool with a common drive. Second, MUs are recruited in a rigid fashion that largely accords with Henneman’s size principle. Although this paradigm has considerable empirical support, a direct test requires simultaneous observations of many MUs across diverse force profiles. In this study, we developed an isometric task that allowed stable MU recordings, in a rhesus macaque, even during rapidly changing forces. Patterns of MU activity were surprisingly behavior-dependent and could be accurately described only by assuming multiple drives. Consistent with flexible descending control, microstimulation of neighboring cortical sites recruited different MUs. Furthermore, the cortical population response displayed sufficient degrees of freedom to potentially exert fine-grained control. Thus, MU activity is flexibly controlled to meet task demands, and cortex may contribute to this ability.
Preprint
Full-text available
Regular exercise promotes whole-body health and prevents disease, yet the underlying molecular mechanisms throughout a whole organism are incompletely understood. Here, the Molecular Transducers of Physical Activity Consortium (MoTrPAC) profiled the temporal transcriptome, proteome, metabolome, lipidome, phosphoproteome, acetylproteome, ubiquitylproteome, epigenome, and immunome in whole blood, plasma, and 18 solid tissues in Rattus norvegicus over 8 weeks of endurance exercise training. The resulting data compendium encompasses 9466 assays across 19 tissues, 25 molecular platforms, and 4 training time points in young adult male and female rats. We identified thousands of shared and tissue- and sex-specific molecular alterations. Temporal multi-omic and multi-tissue analyses demonstrated distinct patterns of tissue remodeling, with widespread regulation of immune, metabolism, heat shock stress response, and mitochondrial pathways. These patterns provide biological insights into the adaptive responses to endurance training over time. For example, exercise training induced heart remodeling via altered activity of the Mef2 family of transcription factors and tyrosine kinases. Translational analyses revealed changes that are consistent with human endurance training data and negatively correlated with disease, including increased phospholipids and decreased triacylglycerols in the liver. Sex differences in training adaptation were widespread, including those in the brain, adrenal gland, lung, and adipose tissue. Integrative analyses generated novel hypotheses of disease relevance, including candidate mechanisms that link training adaptation to non-alcoholic fatty liver disease, inflammatory bowel disease, cardiovascular health, and tissue injury and recovery. The data and analysis results presented in this study will serve as valuable resources for the broader community and are provided in an easily accessible public repository ( https://motrpac-data.org/ ). Highlights Multi-tissue resource identifies 35,439 analytes regulated by endurance exercise training at 5% FDR across 211 combinations of tissues and molecular platforms. Interpretation of systemic and tissue-specific molecular adaptations produced hypotheses to help describe the health benefits induced by exercise. Robust sex-specific responses to endurance exercise training are observed across multiple organs at the molecular level. Deep multi-omic profiling of six tissues defines regulatory signals for tissue adaptation to endurance exercise training. All data are available in a public repository, and processed data, analysis results, and code to reproduce major analyses are additionally available in convenient R packages.
Article
Nuzzo, JL. Narrative review of sex differences in muscle strength, endurance, activation, size, fiber type, and strength training participation rates, preferences, motivations, injuries, and neuromuscular adaptations. J Strength Cond Res 37(2): 494-536, 2023-Biological sex and its relation with exercise participation and sports performance continue to be discussed. Here, the purpose was to inform such discussions by summarizing the literature on sex differences in numerous strength training-related variables and outcomes-muscle strength and endurance, muscle mass and size, muscle fiber type, muscle twitch forces, and voluntary activation; strength training participation rates, motivations, preferences, and practices; and injuries and changes in muscle size and strength with strength training. Male subjects become notably stronger than female subjects around age 15 years. In adults, sex differences in strength are more pronounced in upper-body than lower-body muscles and in concentric than eccentric contractions. Greater male than female strength is not because of higher voluntary activation but to greater muscle mass and type II fiber areas. Men participate in strength training more frequently than women. Men are motivated more by challenge, competition, social recognition, and a desire to increase muscle size and strength. Men also have greater preference for competitive, high-intensity, and upper-body exercise. Women are motivated more by improved attractiveness, muscle "toning," and body mass management. Women have greater preference for supervised and lower-body exercise. Intrasexual competition, mate selection, and the drive for muscularity are likely fundamental causes of exercise behaviors in men and women. Men and women increase muscle size and strength after weeks of strength training, but women experience greater relative strength improvements depending on age and muscle group. Men exhibit higher strength training injury rates. No sex difference exists in strength loss and muscle soreness after muscle-damaging exercise.
Article
Repeated, episodic bouts of skeletal muscle contraction undertaken frequently as structured exercise training is a potent stimulus for physiological adaptation in many organs. Specifically in skeletal muscle, remarkable plasticity is demonstrated by the remodeling of muscle structure and function in terms of muscular size, force, endurance, and contractile velocity as a result of the functional demands induced by various types of exercise training. This plasticity, and the mechanistic basis for adaptations to skeletal muscle in response to exercise training, is underpinned by activation and/or repression of molecular pathways and processes induced in response to each individual acute exercise session. These pathways include the transduction of signals arising from neuronal, mechanical, metabolic, and hormonal stimuli through complex signal transduction networks, which are linked to a myriad of effector proteins involved in the regulation of pre- and post-transcriptional processes, and protein translation and degradation processes. This review therefore describes acute exercise-induced signal transduction and the molecular responses to acute exercise in skeletal muscle including emerging concepts such as epigenetic pre- and post-transcriptional regulation, and the regulation of protein translation and degradation. A critical appraisal of methodological approaches and the current state of knowledge informs a series of recommendations offered as future directions in the field.