Sunteți pe pagina 1din 31

Exercise physiology

1 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

Official reprint from UpToDate


www.uptodate.com 2016 UpToDate

Exercise physiology
Authors
David M Systrom, MD, FRCPC
Gregory D Lewis, MD

Section Editor
James K Stoller, MD, MS

Deputy Editor
Helen Hollingsworth, MD

All topics are updated as new evidence becomes available and our peer review process is complete.
Literature review current through: Dec 2015. | This topic last updated: Jul 22, 2014.
INTRODUCTION Physical exercise requires the coordinated interaction of ventilation, cardiac output, and systemic
and pulmonary blood flow to meet the metabolic demands of contracting muscles, as skeletal muscle metabolism can
rise quickly to 50 times its resting rate during heavy exercise. To preserve cellular oxygenation and acid-base
homeostasis during exercise, metabolic, cardiovascular, respiratory responses must adapt rapidly to these dramatic
changes in tissue demands.
The normal physiologic response to exercise will be reviewed here. The role of exercise testing to evaluate reduced
exercise tolerance due to dysfunction of the respiratory or cardiovascular systems is discussed separately. (See
"Functional exercise testing: Ventilatory gas analysis" and "Exercise ECG testing: Performing the test and interpreting
the ECG results" and "Approach to the patient with dyspnea".)
DEFINITIONS
Minute ventilation Volume of air exhaled per minute, VE
Oxygen uptake (L/min) VO2
Maximum oxygen uptake VO2max
Oxygen consumption in tissues (eg, muscle) QO2
Carbon dioxide output from lungs (L/min) VCO2
Carbon dioxide production in tissues QCO2
Arterial oxygen content CaO2
Mixed venous oxygen content CvO2
Respiratory exchange ratio VCO2/VO2 (measured at the mouth)
Lactate threshold (LT) The level of oxygen uptake in the lungs (VO2) at which a sustained rise in blood lactate occurs
Ventilatory threshold The point during exercise at which the VCO2 increases out of proportion to the VO2, which is
approximately the same point as the LT
Breathing reserve index Ratio of VE at peak exercise (VEmax) to maximal voluntary ventilation (MVV), VEmax/MVV
ORGAN SYSTEM-SPECIFIC ROLES IN EXERCISE PERFORMANCE The integrated function of several systems
(ie, skeletal muscles, energy supply, cardiac output, circulation, respiration) is necessary for a normal response to
physical exercise. The contribution of each of these systems is outlined in the following sections.
Skeletal muscle Skeletal muscles are organized in motor units, composed of between 10 and 2000 muscle fibers.
Each motor unit is innervated by a single motor neuron. The muscle fibers within a given motor unit are classified as
type I or type II based on features such as the oxidative or glycolytic enzyme content, contraction speed, myoglobin
content, and myosin adenosine triphosphatase content (table 1) [1,2]. These features in turn influence the mechanical
output of the muscle unit.
Type I (also called red or slow-twitch) fibers have a high content of the oxygen binding protein myoglobin, a high

06/01/2016 20:05

Exercise physiology

2 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

content of oxidative enzymes for producing adenosine triphosphate (ATP), and a high density of mitochondria [2]. These
fibers tend to be fatigue resistant and are recruited for low-level endurance activity. The high myoglobin content in type
I fibers provides a ready supply of oxygen for oxidative metabolism.
Type II (also called white or fast-twitch) fibers are further categorized as types IIA and IIX (table 1) [2]. Type IIA fibers
are similar to type I in their myoglobin content and oxidative enzyme capacity. Type IIX fibers have a low content of
myoglobin and high anaerobic, glycolytic capacity and are recruited for short-term, heavy work, particularly work above
70 percent of the muscle's aerobic capacity [2].
Fiber type varies considerably among human muscles; as examples, the soleus consists predominantly of type I fibers,
while the triceps is composed mainly of type II fibers, and the vastus lateralis muscle is approximately 50 percent type
I. The fiber type mix within a given muscle varies among individuals and is genetically determined. Training can result in
a greater capillary density, increased muscle fiber size, and an increased concentration of mitochondria within a fiber,
along with the potential for alteration of the proportion of fiber types within a given muscle [2,3]. (See 'Adaptations to
training' below.)
Metabolism in skeletal muscles Muscle contraction and relaxation depend primarily upon hydrolysis of adenosine
triphosphate (ATP), which releases the chemical energy necessary for binding of the protein myosin with actin filaments
to allow myosin to slide along the actin filament leading to mechanical contraction. Energy metabolism in muscle and the
various metabolic myopathies are discussed separately. (See "Energy metabolism in muscle".)
Energy sources The main sources for energy to produce the ATP used by skeletal muscle during exercise are
glycogen, glucose, and free fatty acids with glycogen being the predominant source of energy. Protein is rarely used as
an energy source, except during periods of starvation. The specific energy source used by working muscle for aerobic
metabolism depends upon a number of factors including the intensity, type, and duration of exercise, and also physical
conditioning and diet.
Glycogen, glucose, and free fatty acids all provide energy for the creation of ATP, although the amount of ATP
generated depends on the metabolic pathway used. Hydrolysis of ATP by the enzyme actomyosin-ATPase permits
actin-myosin cross-bridge formation and release, with resultant muscular contraction; adenosine diphosphate (ADP)
and phosphate (PO4-) are released in the process.
Metabolic pathways A number of biochemical processes in muscle fibers are responsible for maintaining a
constant supply of ATP, as intracellular stores of the high-energy compound ATP are small and must be continually
replenished. The three main energy producing pathways that are utilized to prevent significant decreases in ATP
concentration during dynamic exercise are the phosphocreatine shuttle, oxidative phosphorylation, and anaerobic
glycolysis.
Phosphocreatine shuttle Upon initiation of exercise, the first energy "buffer" is the phosphocreatine (PCr)
shuttle, whereby the enzyme creatine kinase splits a molecule of inorganic phosphate (Pi) from PCr, which then
combines with adenosine diphosphate (ADP) to yield ATP and creatine (Cr). The shuttle mechanism maintains the ATP
concentration in the proximity of actin-myosin cross-bridging for short duration exertion [4]:
PCr + ADP > Cr + ATP <> ADP + Pi
When ATP is utilized for muscle contraction, ADP and PO4- are released. Skeletal muscle PCr concentration has been
shown to decrease and recover with incremental exercise [5,6]. Training and oral loading of carbohydrates or creatine
improve PCr kinetics and exercise performance [5,7,8]. (See 'Metabolic system' below.)
Oxidative phosphorylation The most efficient skeletal muscle ATP source is the oxidative phosphorylation
of intracellular glycogen and free fatty acids (FFA) in the muscle mitochondria (figure 1 and figure 2) [9]. In the initial
steps, pyruvate is produced during the metabolism of glycogen, glucose, or FFA and then converted to acetyl
coenzyme A (figure 3). Acetyl coenzyme A enters the tricarboxylic acid (TCA) cycle (also known as the citric acid cycle
or Krebs cycle) with resultant generation of nicotinamide adenine dinucleotide (NADH) and flavin adenine dinucleotide
(FADH2). Oxidative phosphorylation occurs when NADH and FADH2 donate electrons to generate approximately 26

06/01/2016 20:05

Exercise physiology

3 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

molecules of ATP per molecule of pyruvate.


TCA cycle flux in exercising skeletal muscle increases up to 100-fold during strenuous exercise, with associated fivefold
increases in TCA cycle intermediates, known as the intermediate pool [10,11]. Endurance training has been associated
with TCA intermediate pool expansion at the start of exercise and may contribute to improved energy provision [11].
Whether TCA intermediate pool expansion is necessary to support increased flux, thereby rendering the reactions that
refill the TCA cycle critical to TCA cycle function, or whether TCA intermediate pool expansion is simply a pyruvate sink
reflecting glycolytic pyruvate production exceeding the rate of oxidation in the TCA cycle, remains controversial [10].
With more prolonged exercise that depletes the skeletal muscle supplies of carbohydrate and FFA, blood borne
glucose, FFA, and gluconeogenic amino acids (eg, branched chain) become additional sources of energy. The
"gluconeogenic" amino acids can be metabolized to alpha keto acids, and then to glucose, and thereby contribute to
ATP production during exercise, especially when muscle glycogen is depleted [12]. The disadvantage of using amino
acids as an energy source is that they are derived from muscle protein, and protein catabolism eventually leads to loss
of muscle strength.
Anaerobic glycolysis Anaerobic glycolysis is a rapid source of ATP production in which pyruvate derived
from glycolysis (figure 3) is converted to lactate without oxygen, yielding two molecules of ATP. However, anaerobic
glycolysis is less efficient than oxidative phosphorylation as it produces much less ATP than the approximately 26 that
are produced by oxidative phosphorylation [2,13].
The action of adenylate kinase yields the two adenosine diphosphate molecules needed for the reaction:
Glucose + 2Pi + 2ADP <> 2 Lactate + 2H2O + 2ATP
Anaerobic glycolysis results in faster production of ATP than oxidative glycolysis but cannot be sustained due to a rapid
drop in muscle cytoplasmic pH, which inhibits further glycolysis [13]. Once excess lactic acid is moved extracellularly to
the blood, it is buffered by bicarbonate, thus producing additional CO2 (in excess of O2 consumption). Lactate is then
transported to the liver where it is converted to pyruvate, which in turn is metabolized to glucose.
Anaerobic glycolysis is an important ATP source when the capacity of pyruvate dehydrogenase to metabolize pyruvate
through the TCA cycle is exceeded and pyruvate is oxidized via mass action to lactate [14]. These conditions are
fostered by the rapid glycogenolysis associated with elevated plasma catecholamines during heavy, brief exercise and
when there is an imbalance between mitochondrial oxygen supply and demand during ischemic and hypoxic exercise or
at the onset of intense exercise.
Circulatory system An essential role of the circulatory system during exercise is to deliver oxygen to the muscles.
This task is accomplished by increases in extraction of oxygen by exercising muscle, heart rate and stroke volume, and
decreases in systemic and pulmonary vascular resistance.
Fick equation The delivery of oxygen to the muscles by the circulatory system (VO2) is expressed in a
rearrangement of the Fick equation:
VO2 = Qt x (CaO2 - CvO2)
In this equation, Qt is the cardiac output (product of heart rate times stroke volume) and (CaO2 - CvO2) is the systemic
oxygen extraction or the difference in O2 content between arterial and mixed venous blood (figure 4). Increased muscle
activity during exercise results in increased oxygen extraction in the peripheral circulation. (See "Oxygen delivery and
consumption", section on 'Definitions' and "Functional exercise testing: Ventilatory gas analysis", section on 'Aerobic
parameters'.)
The arteriovenous oxygen difference is calculated by the following equation:
(CaO2 - CvO2) = 1.34 hemoglobin concentration (arterial oxygen saturation mixed venous oxygen
saturation)
Maximal cardiac output, which facilitates transport of oxygen from the alveolus to skeletal muscle, determines the

06/01/2016 20:05

Exercise physiology

4 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

maximal oxygen uptake (VO2 max) and aerobic capacity to a large degree. Maintenance of CaO2 and depression of
CvO2 during exercise are also circulatory functions, requiring exquisite matching of blood flow to ventilation and tissue
metabolism, respectively.
Cardiac output Cardiac output increases during incremental exercise through changes in both heart rate (HR)
and stroke volume (SV). The relative increments in each component of cardiac output (HR, SV) and C(a-v)O2 that
permit an increase in VO2 during exercise are illustrated in the figure (figure 5). In healthy adults, cardiac output is
generally the limiting factor in the VO2 max.
Maximal heart rate The maximal HR decreases as a function of age, as predicted by the following equation
[15]:
Maximal HR = 220 - age (in years)
However, a meta-analysis of 18,712 subjects found that this equation underestimates the maximal heart rate in
older subjects [16]. The following equation was more accurate for predicting maximal HR in healthy adults:
Maximal HR = 208 - 0.7 x age (in years)
Normally, the difference between achieved and maximal predicted HR (called the heart rate reserve) is less than
15 beats per minute [17].
The rise in HR is linear versus VO2, initially due to withdrawal of vagal tone, and subsequently due to increased
sympathetic activity [18]. The maximal HR also appears directly related to lean body mass [19]. Thus, the
malnourished or myopathic patient may have a reduced heart rate (versus the normal individual) at peak exercise.
Training results in a lower HR at rest and at any given VO2 (figure 6), but does not affect maximal HR [20]. (See
'Cardiovascular adaptations' below.)
Stroke volume Stroke volume (SV) increases in a hyperbolic fashion versus VO2, and maximum values can be
augmented by up to 100 percent with training [2,21]. The rise in SV during exercise is mediated in part by
increased contractility, reflected by an increase in left ventricular ejection fraction (LVEF) of approximately 10
percent from rest to peak exercise [22]. LV filling is enhanced during exercise by capacitance venoconstriction,
greater negative intrathoracic pressures, and the pumping action of exercising limbs [23]. As a result, left
ventricular end-diastolic volume (LVEDV) also increases by 20 to 40 percent, augmenting SV by the FrankStarling mechanism [24]. (See "Pathophysiology of heart failure: Left ventricular pressure-volume and other
hemodynamic relationships".)
In a heart with normal relaxation (lusitropic) properties, LV end-diastolic pressure increases to approximately 20
mmHg during maximum exercise [25]. Diastolic filling is limited by the physical constraints of the pericardium, as
evidenced by the increase in maximum cardiac output and VO2 following pericardiectomy [26].
Maximum cardiac output Cardiac output normally increases by approximately 5 mL/min for every 1 mL/min
increase in VO2 [27]. This slope is not altered by training, but maximum cardiac output improves with conditioning
to four to five times resting values (up to levels of approximately 25 L/min in a young healthy individual). Maximum
cardiac outputs above 40 L/min have been reported in elite athletes, and elite athletes may exhaust their breathing
reserve before attaining maximal cardiac output [28].
Abnormal cardiovascular reserve function Exercise-induced elevations in left sided filling pressures (ie,
exercise PCWP >20 mmHg) can unmask heart failure with preserved ejection fraction (HFpEF). Cardiac
contractility may be relatively preserved at rest in HFpEF, despite dramatic limitations in the ability to augment
systolic and vascular function with the stress of exercise. This is termed abnormal cardiovascular reserve function
[29]. Abnormal endothelium-dependent vasodilation contributes importantly to vascular stiffening and to abnormal
flow-dependent vasodilation with exercise, and aortic stiffness has been most strongly related to abnormal
exercise capacity in HFpEF patients [30,31]. In a small cohort of patients with HFpEF, exercise capacity was
markedly reduced, and this was related to an inability to increase cardiac output by enhancing preload (end
diastolic volume) [32]. However, additional invasive hemodynamic studies are needed to precisely define the

06/01/2016 20:05

Exercise physiology

5 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

functional consequences of elevated left-sided filling pressures during exercise. Pulmonary capillary wedge
pressure normally increases about 1.4 mmHg for every 1 mmHg increase in right atrial pressure, suggesting
interdependence of right and left ventricular filling [33]. (See "Clinical manifestations and diagnosis of heart failure
with preserved ejection fraction".)
Systemic circulation Several changes occur in the systemic circulation to enhance oxygen delivery to the
exercising muscles. Systolic blood pressure (BP) increases via the muscle chemoreflex during exercise, but diastolic
pressure generally remains near resting values [23,34,35]. The rise in mean systemic BP with exertion is normally much
less than the increment in cardiac output, reflecting a decrease in systemic vascular resistance (SVR) [34]. The ability
to decrease SVR has been correlated with exercise performance [36].
Blood flow during exercise is preferentially directed to working muscle and away from less metabolically active tissues
such as gut and kidney (table 2) [27]. Distribution of blood flow is influenced by sympathetic arterial vasoconstriction
and skeletal muscle vasodilation due to local decreases in pH and PaO2, and local increases in potassium, adenosine,
and nitric oxide (NO) [37,38]. Systemic vasodilation by endogenous nitric oxide (NO) is enhanced by training [39].
Normal peripheral vasoregulation and a rightward shift (decreased oxygen affinity) of the oxyhemoglobin dissociation
curve with acidosis increase skeletal muscle oxygen extraction, yielding a femoral venous PO2 near 20 mmHg and a
systemic O2 extraction ratio ([CaO2 - CvO2] CaO2) as high as 0.75 (figure 7) [40-42]. By comparison, a normal
resting systemic O2 extraction rate is approximately 0.25 to 0.30. (See "Oxygen delivery and consumption", section on
'Oxygen extraction'.)
Pulmonary circulation The pulmonary circulation normally receives >95 percent of the cardiac output, and does
so with minimal resistance [24]. Similar to the systemic circulation, the pressure gradient across the pulmonary vascular
bed increases during exercise by a smaller factor than the increase in blood flow due to a fall in pulmonary vascular
resistance (PVR) from approximately 96 to 60 dynes-sec-cm-5 [26]. The decrease in PVR during exercise is a
consequence of passive distention of a compliant circulation, active vasodilation mediated in part by NO, and an
increase in cardiac output by up to five times baseline [43].
Exercise pulmonary hemodynamic results have been reported in 218 normal subjects [44]. However, establishment of
"normal" hemodynamic values across age and sex strata for pulmonary arterial pressure and pulmonary capillary
wedge pressure requires further investigation. Nonetheless, in normal subjects, the increase in mean pulmonary artery
pressure during exercise should not exceed 3 mmHg per liter of increase in cardiac output [45].
In a study of 406 subjects undergoing cardiopulmonary exercise testing (CPET) for evaluation of dyspnea, continuous
hemodynamic monitoring demonstrated that exercise-induced pulmonary hypertension (EIPH), defined as an exercise
mean pulmonary artery pressure (PAP) >30 mmHg, a pulmonary capillary wedge pressure <20 mmHg, and failure of
pulmonary vascular resistance (PVR) to fall below 80 dynes-sec-cm-5, was associated with poor exercise tolerance
[46]. These data support the hypothesis that adequate pulmonary vasodilation is critical to enable the thin walled right
ventricle to augment cardiac output during exercise. Exercise-related pulmonary hypertension may unmask the early
diagnosis of pulmonary arterial hypertension. (See "Clinical features and diagnosis of pulmonary hypertension in adults",
section on 'Diagnostic evaluation'.)
Respiratory system One of the most remarkable aspects of exercise physiology is the maintenance of arterial
oxygen and carbon dioxide levels (PaO2, PaCO2) within narrow ranges in the face of the large, rapid increases in
metabolic rate that characterize vigorous exertion. These indices are preserved largely because of adaptations in
ventilation and ventilation/perfusion matching.
Ventilation The minute ventilation (VE), defined as the volume of air that is exhaled (or inhaled) in one minute,
normally rises during incremental exercise as a result of a linear increase in breathing frequency (up to about 50
breaths/minute in normal adults) and a hyperbolic increase in tidal volume (VT, the volume of air inhaled per breath)
[47]. VT reaches a plateau at approximately 50 percent of the resting vital capacity, above which the elastic work of
breathing is prohibitive [48]. Overall, VE can increase approximately 10-fold with intense exertion.
Regulation of ventilation The mechanisms by which VE is regulated during exercise are incompletely

06/01/2016 20:05

Exercise physiology

6 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

understood. As an example, the increase in VE during submaximal exercise perfectly matches carbon dioxide
production without any detectable antecedent changes in arterial pH or PaCO2 to initiate the process. Lactic acidemic
stimulation of the carotid bodies has been proposed to drive ventilation during intense exercise, but patients with
McArdle's syndrome regulate ventilation normally during exercise despite minimal lactate production [49]. The
observation that denervation of the carotid bodies in dogs actually increases the exercise ventilatory response provides
further evidence that the chemoreceptors may not be necessary for the hyperpnea of heavy exercise [50]. (See
"Myophosphorylase deficiency (glycogen storage disease V, McArdle disease)".)
Functional magnetic resonance imaging studies of the human central nervous system suggest that stimuli for autonomic
ventilatory control are integrated in the ventrolateral medulla [51]. A feed-forward locomotor-linked mechanism
emanating from the hypothalamus is capable of stimulating both ventilation and exercise motion in parallel [52]. In
addition, a muscle chemoreflex, stimulated by the local accumulation of metabolic products of exercise, can also drive
ventilation during exercise [53,54]. (See "Control of ventilation".)
Ventilatory threshold The ventilatory threshold is defined as the point at which the VE increases out of
proportion to the VO2. The ventilatory threshold occurs at approximately the same time as the patient approaches the
lactate threshold (LT, the level of VO2 at which a sustained rise in blood lactate occurs) and metabolic acidemia
ensues. The ventilatory threshold can be used as a noninvasive marker of the LT because these two events normally
occur at a nearly identical level of exercise [55]. (See 'Lactate threshold' below.)
The ventilatory threshold is best demonstrated by plotting VE/VO2 and VE/VCO2 on a graph versus exercise intensity
or stage, where VCO2 is the volume of CO2 produced (L/min). The ventilatory threshold is the point just before
VE/VO2 begins to rise without a similar rise in VE/VCO2 (figure 8).
Breathing reserve index The ratio of VE at peak exercise (VEmax) to the maximal voluntary ventilation
(MVV) at rest has been termed the breathing reserve index (BRI) [56]. To calculate the BRI, the maximal voluntary
ventilation can either be measured by a 12-second voluntary effort or estimated by multiplying the FEV1 by 40 [57-59].
(See "Overview of pulmonary function testing in adults".)
A BRI (VEmax/MVV) of 0.70 can be sustained for 15 minutes in normal individuals, but values above 0.75 are usually
not attained even at peak exercise in untrained individuals [60]. This observation suggests that VO2 max in normal
individuals is limited by cardiac factors, not ventilation [61,62]. Elite athletes may reach such a high level of
cardiovascular fitness that a pulmonary mechanical limitation to exercise is approached, but this is distinctly unusual
[28].
Patients with intrinsic lung disease typically have a low BRI.
Carbon dioxide elimination For normal individuals, ventilation is regulated to maintain the arterial partial
pressure of carbon dioxide (PaCO2) at isocapnic (constant PaCO2) levels by increasing or decreasing VCO2. When
exercise intensity surpasses aerobic capacity leading to anaerobic glycolysis, respiratory compensation for metabolic
acidosis occurs, and the increased VCO2 results in a decrease in PaCO2 [47,63]. Maximum exercise produces a
partially compensated metabolic acidosis, with an arterial pH of 7.20 to 7.30 [64,65], while short-term exercise to
exhaustion can reduce the arterial pH to 7.15 or less [66]. (See 'Anaerobic glycolysis' above.)
Elimination of carbon dioxide in the lungs (VCO2) is achieved by increased alveolar ventilation. Alveolar ventilation
reflects the component of tidal volume (VT) that participates in gas exchange; dead-space ventilation (VD) refers to
that component of VT that helps move air to and from alveoli, but does not participate in gas exchange. Increased
alveolar ventilation is associated with a decrease in VD/VT.
Anatomic dead space increases during exercise because of a tethering effect on conducting airways at high VT,
whereas alveolar dead space decreases because of augmented blood flow to the lung apices, giving the net effect of a
slight increase in total (physiologic) VD. However, this effect is more than offset by the increased VT, which produces a
decrease in upright VD/VT from up to 0.45 at rest to less than 0.29 at maximum exercise [2].
The relationship between ventilation and CO2 elimination during exercise is described by the alveolar ventilation

06/01/2016 20:05

Exercise physiology

7 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

equation:
VE = (863 x VCO2)/PaCO2 (1-VD/VT)
From this equation, the amount of ventilation required for exercise is defined by three factors:
Carbon dioxide output (VCO2)
The set point at which PaCO2 is regulated by ventilatory control mechanisms
The ratio of physiologic dead-space (VD) to tidal volume (VT), or VD/VT
The amount of ventilation required for the lungs to eliminate CO2, expressed as the ratio of VE/VCO2 at the anaerobic
threshold, or the slope of VE/VCO2 during exercise, has been termed "ventilatory efficiency." Impaired ventilatory
efficiency (ie, VE/VCO2 slope >28 or VE/VCO2 at anaerobic threshold [AT] >36) is increasingly recognized as an
important determinant of prognosis in disease states associated with increased VD/VT such as pulmonary arterial
hypertension and heart failure (figure 9) [67].
In normal individuals, VE/VCO2 does not differ between genders but increases with age. CO2 elimination by the lungs
becomes more efficient during exercise [63].
Oxygenation During exercise, the partial pressure of oxygen in arterial blood (PaO2) remains near resting
values despite marked reductions in mixed venous oxygen tension (PvO2) and an abbreviated red cell transit time
through the pulmonary capillaries. Arterial oxygenation is maintained by several adaptations, including increased
alveolar oxygen tension (PAO2), a decreased number of low V/Q units, increased surface area for O2 diffusion, and a
smaller right-to-left shunt fraction.
The driving pressure for O2 diffusion across the alveolar-capillary membrane is the PAO2, as described by the
equation:
PAO2 = PiO2 - (PACO2 RER)
where PiO2 is the inspired partial pressure of O2 and RER is the respiratory exchange ratio (VO2/VCO2).
PAO2, which is roughly equal to PaCO2, increases in normal persons exercising above the lactate threshold at sea level
because of hyperventilation (which lowers PACO2) and an increased RER, due to bicarbonate buffering of lactic acid
and "non-metabolic" CO2 production. The excess CO2 is often referred to as "non-metabolic" because it is not
produced by metabolism, per se.
Oxygenation also improves in the lung with exercise because of a decrease in the number of low-ventilation/perfusion
units at the bases due to larger tidal breaths. Improved distribution of pulmonary blood flow also results in augmentation
of the diffusing surface area at high cardiac outputs. Finally, in a manner analogous to VD/VT, the right-to-left shunt
fraction (Qs/Qt) falls because of the large increase in Qt.
Some well-trained individuals manifest arterial O2 desaturation at extremely high metabolic rates. This has been
attributed to reduced compensatory hyperventilation and a diffusion limitation resulting from rapid red cell transit time
through the pulmonary capillaries [68,69].
Hematologic contribution The oxygen carrying capacity of blood (amount of oxygen bound to hemoglobin plus the
amount of oxygen dissolved in arterial blood) is a significant contributor to exercise capacity. Therefore, hemoglobin
level should be taken into account when interpreting cardiopulmonary exercise test results. (See "Oxygen delivery and
consumption".)
ASSESSMENT OF EXERCISE CAPACITY Cardiopulmonary exercise testing (also known as multiparameter
exercise testing) yields detailed information on an individual's response to exercise [2,70]. The symptom-limited,
incremental exercise test involves a continuous, ramped increase in workload that continues until the patient has
symptoms (eg, dyspnea, fatigue) that cause the patient to feel unable to exercise at a higher workload. Physiologic
data, including oxygen uptake, carbon dioxide output, tidal volume, minute ventilation, electrocardiographic (ECG)

06/01/2016 20:05

Exercise physiology

8 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

tracings, and pulse oximetry are measured throughout the test and during the first several minutes of recovery. (See
"Functional exercise testing: Ventilatory gas analysis" and "Exercise capacity and VO2 in heart failure".)
Maximum oxygen uptake The maximum oxygen uptake (VO2 max, L/minute) reflects the maximal ability of a
person to take in, transport, and use oxygen, and it defines that person's functional aerobic capacity. VO2 max has
become the "gold standard" laboratory measure of cardiorespiratory fitness and is the most important parameter
measured during functional exercise testing. The VO2 max attained during incremental bilateral leg exercise (eg, bicycle
ergometer, treadmill) is used to provide an overall assessment of exercise capacity [71-74]. VO2 increases linearly
versus work rate with a slope of approximately 10 mL/min per watt in normal subjects [64]. This slope is not affected
by age, sex, or training but is shifted leftward in obese patients (figure 10).
A true VO2 max is identified by a plateau of VO2 when VO2 is graphed versus work, but this plateau occurs only in a
subset of normal subjects and patients, usually during a maximal incremental protocol [74]. In cases where VO2 max is
not achieved, peak VO2, averaged over 30 of the final 60 seconds of exercise, is the more appropriate term to
describe the highest achieved oxygen uptake during exercise.
VO2 max is often indexed to body weight (mL/kg per min) or expressed in metabolic equivalents (METS), which are
multiples of normal baseline oxygen uptake at rest. One metabolic equivalent (MET) is equal to 3.5 mL/kg/min. Normal
reference values have been reported from a population-based survey in which rigorous phenotyping was used to limit
analyses to healthy individuals (table 3) [75]. These values replace previous values derived from male shipyard workers
[64], university members [76], and soldiers [77].
Of note, VO2 max achieved during cycle ergometry, as illustrated in the table (table 3), is 5 to 11 percent lower than
that achieved during exercise treadmill testing, due to lower muscle mass utilized during cycle ergometry [78,79]. The
normal VO2 max can be predicted from age, gender, height, and lean body weight.
A normal VO2 max usually exonerates the pulmonary, cardiovascular, and neuromuscular systems of serious pathology,
although intra- or inter-organ compensation for mild primary abnormalities can result in a relatively normal value. As an
example, patients with chronotropic disturbances of the heart provide an example of intra-organ compensation;
near-normal values for cardiac output may be preserved during exercise by relative increases in stroke volume, via the
Frank-Starling mechanism [71]. Examples of inter-organ compensation include the elevation in cardiac output commonly
seen in anemic individuals [80], and the improvement in systemic O2 extraction which occurs in some patients with
cardiac dysfunction [81].
Maximum work The maximum work (in watts; 1 watt equals 0.0143 kcal/min) achieved during an incremental
exercise test has been used to evaluate overall exercise capacity. However, this variable can be misleading because a
significant amount of work can be performed by obese patients and those with obstructive airways disease beyond that
which is measured by the cycle ergometer or treadmill [70,82]. As an example, the work of breathing in obstructive
airways disease may be substantial. In such a situation, a spuriously low maximal external work rate may be recorded
even though the more meaningful value of VO2 max is relatively normal.
Lactate threshold Heavy workloads are associated with an increase in blood lactate concentration, although the
specific workload that results in an increase in lactate concentration varies from one individual to another (figure 11).
The level of oxygen uptake in the lungs (VO2) at which a sustained rise in blood lactate occurs is called the lactate
threshold (LT). The anaerobic threshold is the level of VO2 above which anaerobically produced ATP supplements
aerobic ATP production, and both the blood level of lactate and the ratio of lactate to pyruvate increase. Practically
speaking these metabolic events occur together. (See 'Anaerobic glycolysis' above.)
For many years the lactic acidemia of exercise was assumed to be secondary to inadequate oxygen (O2) delivery to
muscle with resultant increases in anaerobic glycolysis to produce ATP. However, since the 1980s, it has been
appreciated that the skeletal muscle mitochondrial redox state is actually higher when working muscle is producing
lactate than at rest, implying that oxygen supply is not the critical factor [14,15,83]. Currently, the LT is considered to
be the VO2 at which pyruvate, and therefore lactate, production exceeds its ability to be metabolized via the TCA cycle
[14].

06/01/2016 20:05

Exercise physiology

9 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

At a VO2 just below LT, steady state metabolic conditions remain preserved, and exercise can be sustained for
prolonged periods [16]. The LT varies with cardiovascular fitness and is a useful clinical index [17,18].
The LT occurs at greater than 40 percent of the predicted VO2 max in normal individuals but earlier in the course of
exercise among patients with cardiovascular disease [19,20]. (See "Exercise capacity and VO2 in heart failure".)
Exercise intensity must approach the LT for optimal training effects to occur [84]. An endurance athlete will not reach an
LT until 80 to 90 percent of the VO2 max and usually competes at a metabolic rate just below the LT.
Fatigue Fatigue during exercise has both central (nervous system) and peripheral (muscle) components, of which
the latter is better understood [85,86]. The development of peripheral fatigue depends upon the intensity and duration
of exercise and is influenced by:
Accumulation of metabolic byproducts
Depletion of high-energy phosphates
Depletion of glycogen substrate
During brief, intense exercise, skeletal muscle intracellular pH decreases because of the accumulation of cytosolic
lactate and loss of potassium [87]. Increased concentration of hydrogen ion in the myocyte may contribute to fatigue
through inhibition of glucosyl flux and muscle contraction [66,88]. With acute exercise to exhaustion, for example, the
intracellular lactate concentration can exceed 40 meq/L and the pH may fall below 6.40 [89].
Ammonia, generated from deamination of AMP to inosine monophosphate in the stressed type II muscle fiber, has also
been implicated in peripheral fatigue, possibly via inhibition of oxidative phosphorylation [90]. Training is associated with
relatively less ammonia and lactate in muscle and blood during exercise and with less fatigue at a given metabolic rate
[86,91].
Depletion of high-energy phosphate compounds also probably plays a role in peripheral fatigue with short-term exercise
[5]. Evidence for the importance of this phenomenon is provided by patients with McArdle's disease, who are incapable
of producing significant quantities of lactate but who fatigue quickly during exercise [92]. In contrast, fatigue during
endurance events appears related to the depletion of glycogen stores rather than high-energy phosphates. This is
thought to be responsible for the phenomenon of the marathon runner "hitting the wall." Muscle metabolism rarely is the
critical factor in determining maximal exercise tolerance, but it may affect an individual's ability to sustain high levels of
exertion [83]. (See 'Metabolic system' below.)
ADAPTATIONS TO TRAINING Long-term adaptations to exercise training include effects upon the musculoskeletal,
metabolic, cardiovascular, and respiratory systems. The improvements in muscle and cardiorespiratory function with
endurance training increase the maximal oxygen uptake (VO2 max) and the lactate threshold [93,94]. Thus, the
endurance trained individual can perform at higher rates of work than an untrained person. (See 'Maximum oxygen
uptake' above and 'Ventilatory threshold' above.)
Musculoskeletal system Skeletal muscle adapts to regular physical activity training with a variety of changes [2,3].
The type of training (eg, prolonged endurance or resistance) affects the type of muscular adaptations [3]. Endurance
training leads to mitochondrial biogenesis, fast-to-slow fiber transformation, expansion of the muscle capillary bed, and
changes in substrate metabolism. Resistance training typically increases the size of muscle fibers, which leads to the
ability to exert more force [3]. Prolonged and high-intensity exercise has been shown to increase strength and the cross
sectional area of ligaments and tendons. In general, women and men of all ages show gains in strength from resistance
training, although the degree of adaptation to training varies from one individual to another [95]. Some individuals have
virtually no gain in muscle mass with resistance training, while others experience up to a 60 percent increase [95].
Training-related expansion of the muscle capillary bed allows greater blood flow to active muscles and a more efficient
delivery of oxygen and energy sources.
Metabolic system Metabolic adaptations to endurance training include the following [2,96,97]:
Increase in the size and number of muscle mitochondria

06/01/2016 20:05

Exercise physiology

10 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

Increase in the capacity of skeletal muscle to store glycogen


Expansion in pool of tricarboxylic cycle intermediates in muscle mitochondria (see 'Oxidative phosphorylation'
above)
Improvement in fat utilization as an energy source by trained muscles, which spares glycogen stores [98]
Cardiovascular adaptations Important changes occur in the cardiovascular system in response to endurance
training [99-101]:
Cardiac muscle fibers hypertrophy, the muscle mass of the ventricles increases, and the force of contraction is
greater.
Well-trained athletes have substantial enlargement of the left ventricle (LV) to a degree compatible with a primary
dilated cardiomyopathy. However, global left ventricular systolic function is normal and without regional wall motion
abnormalities.
The total peripheral resistance decreases due to enhanced capillary capacity for blood flow, which improves
oxygen and nutrient delivery to exercising muscles.
The increases in force of myocardial contraction and LV size together enable an increase in stroke volume.
Respiratory adaptations One of the respiratory adaptations to exercise training is a decrease in minute ventilation
(VE) to achieve a given VO2 or VCO2, although the effects of training are most pronounced at levels of exercise beyond
the lactate threshold. The lower VE observed after training appears to be a function of ventilatory training rather than
changes in the control of respiration. (See 'Respiratory system' above.)
The maximal oxygen uptake during exercise (VO2 max) increases as a function of training [67]. A young, world-class
endurance athlete may have a VO2 max greater than 80 mL/kg per minute. Historically, values in women have been
considered less than those of age-matched men of similar size, but a significant portion of the difference has been
shown to be related to training status [9]. Biomechanical efficiency for running and cycling may be less for women, but
during certain endurance events, enhanced oxidation of fat by women is glycogen-sparing and may be advantageous
[9].
Lactate levels at VO2 max are lower for a given level of work in trained athletes than in sedentary controls. In addition,
endurance training of normal subjects leads to a decrease in blood lactate for a given level of exertion compared with
values prior to training.
SUMMARY
Skeletal muscle metabolism can rise quickly up to 50 times its resting rate during heavy exercise. This is
accomplished by increases in minute ventilation, cardiac output, and systemic oxygen extraction along with
microvascular adaptations that increase the delivery of oxygen to the muscles. (See 'Introduction' above.)
Human skeletal muscles are composed of type I (also called red or slow twitch) fibers and type II (also called
white or fast twitch) fibers (table 1). Type II fibers are further categorized as types IIA and IIX. For each
individual, the relative proportion of fiber types influences the capacity for the particular type of exercise (eg,
low-level endurance, rapid heavy work). (See 'Skeletal muscle' above.)
The main energy sources used by skeletal muscles during exercise are glycogen, glucose, and free fatty acids.
Protein is rarely used as an energy source, except during periods of starvation. These substrates provide the
chemical energy to create adenosine triphosphate (ATP), which is used for myosin cross-linking to actin. (See
'Energy sources' above.)
Upon initiation of exercise, the most immediate source of energy is the phosphocreatine (PCr) shuttle, in which the
enzyme creatine kinase splits a phosphate molecule off PCr, and the phosphate molecule then combines with
adenosine diphosphate to make ATP. ATP is then immediately available to myosin for contraction. (See
'Phosphocreatine shuttle' above.)

06/01/2016 20:05

Exercise physiology

11 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

The most efficient source of ATP comes from metabolism glucose and glycogen to pyruvate by glycolysis
followed by metabolism of pyruvate via the tricarboxylic acid (TCA) cycle (also known as the citric acid cycle or
Krebs cycle) in the mitochondria. During low-level exercise, the majority of muscle metabolism is aerobic, meaning
that there is adequate oxygen to metabolize pyruvate by oxidative phosphorylation which provides the maximum
amount of ATP. Alternatively, fatty acids or, rarely, protein are metabolized to acetic acid and enter the TCA cycle.
(See 'Energy sources' above and 'Oxidative phosphorylation' above.)
With progressively increasing muscular work, the capacity of pyruvate dehydrogenase to metabolize pyruvate is
exceeded, and anaerobic glycolysis becomes the source of ATP (figure 2). This point is known as the lactate
threshold (LT), or the level of oxygen uptake in the lungs (VO2) at which a sustained rise in blood lactate occurs
(figure 11). Anaerobic glycolysis metabolizes pyruvate to lactate with a lower yield of ATP than the TCA cycle.
(See 'Metabolic pathways' above.)
The ventilatory threshold is defined as the point at which the minute ventilation (VE) increases out of proportion to
the VO2 (figure 8) and occurs at approximately the same time as the patient approaches the lactate threshold
(LT). Thus, the ventilatory threshold is sometimes used as a noninvasive marker of the LT. (See 'Lactate
threshold' above and 'Ventilatory threshold' above.)
The maximal cardiac output normally sets the limit on aerobic exercise capacity, although endurance training
typically leads to increased cardiac output. (See 'Circulatory system' above.)
The maximum oxygen uptake (VO2 max) reflects the maximal ability of a person to take in, transport, and use
oxygen, and it defines that person's functional aerobic capacity (figure 10). The VO2 max has become the "gold
standard" laboratory measure of cardiorespiratory fitness and is the most important parameter measured during
functional exercise testing. (See 'Assessment of exercise capacity' above.)
The LT occurs at above 40 percent of the predicted VO2 max in normal individuals. In comparison, LT occurs
earlier (at a lower percent of VO2 max) in the course of exercise among patients with cardiovascular disease, but
later (at a higher percent of VO2 max) in endurance trained athletes. (See 'Definitions' above and 'Lactate
threshold' above.)
Training enhances virtually every step of exercise gas exchange from the lung to the skeletal muscle
mitochondrion. Endurance training leads to mitochondrial biogenesis, fast-to-slow fiber transformation, changes in
substrate metabolism, expansion of the muscle capillary bed, and increased cardiac output. Resistance training
typically increases the size and protein content of muscle fibers, which leads to the ability to exert more force.
(See 'Adaptations to training' above.)
The clinical use of cardiopulmonary exercise testing is discussed separately. (See "Functional exercise testing:
Ventilatory gas analysis" and "Evaluation of pulmonary disability" and "Preoperative evaluation for lung resection"
and "Exercise capacity and VO2 in heart failure".)
Use of UpToDate is subject to the Subscription and License Agreement.
REFERENCES
1. Fry AC, Allemeier CA, Staron RS. Correlation between percentage fiber type area and myosin heavy chain
content in human skeletal muscle. Eur J Appl Physiol Occup Physiol 1994; 68:246.
2. Wasserman K, Hansen JE, Sue DY, et al. Physiology of exercise. In: Principles of exercise testing and
interpretation: Including pathophysiology and clinical applications, 5th, Wolters Kluwer Lippincott Williams
&Wilkins, 2011.
3. Coffey VG, Hawley JA. The molecular bases of training adaptation. Sports Med 2007; 37:737.
4. Meyer RA, Sweeney HL, Kushmerick MJ. A simple analysis of the "phosphocreatine shuttle". Am J Physiol 1984;
246:C365.
5. Larson DE, Hesslink RL, Hrovat MI, et al. Dietary effects on exercising muscle metabolism and performance by

06/01/2016 20:05

Exercise physiology

12 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

31P-MRS. J Appl Physiol (1985) 1994; 77:1108.


6. Thompson CH, Kemp GJ, Sanderson AL, Radda GK. Skeletal muscle mitochondrial function studied by kinetic
analysis of postexercise phosphocreatine resynthesis. J Appl Physiol (1985) 1995; 78:2131.
7. Minotti JR, Johnson EC, Hudson TL, et al. Forearm metabolic asymmetry detected by 31P-NMR during
submaximal exercise. J Appl Physiol (1985) 1989; 67:324.
8. Maughan RJ, Greenhaff PL, Leiper JB, et al. Diet composition and the performance of high-intensity exercise. J
Sports Sci 1997; 15:265.
9. Whipp BJ, Ward SA. Will women soon outrun men? Nature 1992; 355:25.
10. Gibala MJ, MacLean DA, Graham TE, Saltin B. Tricarboxylic acid cycle intermediate pool size and estimated
cycle flux in human muscle during exercise. Am J Physiol 1998; 275:E235.
11. Bowtell JL, Marwood S, Bruce M, et al. Tricarboxylic acid cycle intermediate pool size: functional importance for
oxidative metabolism in exercising human skeletal muscle. Sports Med 2007; 37:1071.
12. MacLean DA, Graham TE, Saltin B. Stimulation of muscle ammonia production during exercise following
branched-chain amino acid supplementation in humans. J Physiol 1996; 493 ( Pt 3):909.
13. Kemp GJ, Sanderson AL, Thompson CH, Radda GK. Regulation of oxidative and glycogenolytic ATP synthesis in
exercising rat skeletal muscle studied by 31P magnetic resonance spectroscopy. NMR Biomed 1996; 9:261.
14. Putman CT, Spriet LL, Hultman E, et al. Skeletal muscle pyruvate dehydrogenase activity during acetate infusion
in humans. Am J Physiol 1995; 268:E1007.
15. Graham TE, Saltin B. Estimation of the mitochondrial redox state in human skeletal muscle during exercise. J
Appl Physiol (1985) 1989; 66:561.
16. Whipp BJ, Wasserman K. Oxygen uptake kinetics for various intensities of constant-load work. J Appl Physiol
1972; 33:351.
17. Poole DC, Gaesser GA. Response of ventilatory and lactate thresholds to continuous and interval training. J Appl
Physiol (1985) 1985; 58:1115.
18. Gaesser GA, Poole DC. Lactate and ventilatory thresholds: disparity in time course of adaptations to training. J
Appl Physiol (1985) 1986; 61:999.
19. Sue DY, Hansen JE. Normal values in adults during exercise testing. Clin Chest Med 1984; 5:89.
20. WASSERMAN K, MCILROY MB. DETECTING THE THRESHOLD OF ANAEROBIC METABOLISM IN CARDIAC
PATIENTS DURING EXERCISE. Am J Cardiol 1964; 14:844.
21. Blomqvist CG, Saltin B. Cardiovascular adaptations to physical training. Annu Rev Physiol 1983; 45:169.
22. Mahler DA, Matthay RA, Snyder PE, et al. Volumetric responses of right and left ventricles during upright
exercise in normal subjects. J Appl Physiol (1985) 1985; 58:1818.
23. Shepherd JT. Circulatory response to exercise in health. Circulation 1987; 76:VI3.
24. Manyari DE, Kostuk WJ. Left and right ventricular function at rest and during bicycle exercise in the supine and
sitting positions in normal subjects and patients with coronary artery disease. Assessment by radionuclide
ventriculography. Am J Cardiol 1983; 51:36.
25. Reeves JT, Moon RE, Grover RF, Groves BM. Increased wedge pressure facilitates decreased lung vascular
resistance during upright exercise. Chest 1988; 93:97S.
26. Stray-Gundersen J, Musch TI, Haidet GC, et al. The effect of pericardiectomy on maximal oxygen consumption
and maximal cardiac output in untrained dogs. Circ Res 1986; 58:523.
27. Wade O, Bishop J, eds. Cardiac Output and Regional Blood Flow: FA Davis, 1962.
28. Dempsey JA. J.B. Wolffe memorial lecture. Is the lung built for exercise? Med Sci Sports Exerc 1986; 18:143.
29. Borlaug BA, Melenovsky V, Russell SD, et al. Impaired chronotropic and vasodilator reserves limit exercise
capacity in patients with heart failure and a preserved ejection fraction. Circulation 2006; 114:2138.
30. Kawaguchi M, Hay I, Fetics B, Kass DA. Combined ventricular systolic and arterial stiffening in patients with heart
failure and preserved ejection fraction: implications for systolic and diastolic reserve limitations. Circulation 2003;
107:714.
31. Borlaug BA, Melenovsky V, Redfield MM, et al. Impact of arterial load and loading sequence on left ventricular
tissue velocities in humans. J Am Coll Cardiol 2007; 50:1570.
32. Kitzman DW, Higginbotham MB, Cobb FR, et al. Exercise intolerance in patients with heart failure and preserved

06/01/2016 20:05

Exercise physiology

13 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

left ventricular systolic function: failure of the Frank-Starling mechanism. J Am Coll Cardiol 1991; 17:1065.
33. Reeves JT, Groves BM, Cymerman A, et al. Operation Everest II: cardiac filling pressures during cycle exercise
at sea level. Respir Physiol 1990; 80:147.
34. Francis GS. Hemodynamic and neurohumoral responses to dynamic exercise: normal subjects versus patients
with heart disease. Circulation 1987; 76:VI11.
35. Rowell LB, O'Leary DS. Reflex control of the circulation during exercise: chemoreflexes and mechanoreflexes. J
Appl Physiol (1985) 1990; 69:407.
36. Clausen JP. Circulatory adjustments to dynamic exercise and effect of physical training in normal subjects and in
patients with coronary artery disease. Prog Cardiovasc Dis 1976; 18:459.
37. Hickner RC, Fisher JS, Ehsani AA, Kohrt WM. Role of nitric oxide in skeletal muscle blood flow at rest and during
dynamic exercise in humans. Am J Physiol 1997; 273:H405.
38. Skinner NS Jr, Costin JC. Role of O2 and K+ in abolition of sympathetic vasoconstriction in dog skeletal muscle.
Am J Physiol 1969; 217:438.
39. Koller A, Huang A, Sun D, Kaley G. Exercise training augments flow-dependent dilation in rat skeletal muscle
arterioles. Role of endothelial nitric oxide and prostaglandins. Circ Res 1995; 76:544.
40. Stringer W, Wasserman K, Casaburi R, et al. Lactic acidosis as a facilitator of oxyhemoglobin dissociation during
exercise. J Appl Physiol (1985) 1994; 76:1462.
41. Roca J, Hogan MC, Story D, et al. Evidence for tissue diffusion limitation of VO2max in normal humans. J Appl
Physiol (1985) 1989; 67:291.
42. Li Y, Dash RK, Kim J, et al. Role of NADH/NAD+ transport activity and glycogen store on skeletal muscle energy
metabolism during exercise: in silico studies. Am J Physiol Cell Physiol 2009; 296:C25.
43. Kane DW, Tesauro T, Koizumi T, et al. Exercise-induced pulmonary vasoconstriction during combined blockade of
nitric oxide synthase and beta adrenergic receptors. J Clin Invest 1994; 93:677.
44. Champion HC, Michelakis ED, Hassoun PM. Comprehensive invasive and noninvasive approach to the right
ventricle-pulmonary circulation unit: state of the art and clinical and research implications. Circulation 2009;
120:992.
45. Lewis GD, Bossone E, Naeije R, et al. Pulmonary vascular hemodynamic response to exercise in
cardiopulmonary diseases. Circulation 2013; 128:1470.
46. Tolle JJ, Waxman AB, Van Horn TL, et al. Exercise-induced pulmonary arterial hypertension. Circulation 2008;
118:2183.
47. Wasserman K, Whipp BJ. Excercise physiology in health and disease. Am Rev Respir Dis 1975; 112:219.
48. Hey EN, Lloyd BB, Cunningham DJ, et al. Effects of various respiratory stimuli on the depth and frequency of
breathing in man. Respir Physiol 1966; 1:193.
49. Hagberg JM, King DS, Rogers MA, et al. Exercise and recovery ventilatory and VO2 responses of patients with
McArdle's disease. J Appl Physiol (1985) 1990; 68:1393.
50. Dempsey JA. Exercise hyperpnea. Chairman's introduction. Adv Exp Med Biol 1995; 393:133.
51. Gozal D, Hathout GM, Kirlew KA, et al. Localization of putative neural respiratory regions in the human by
functional magnetic resonance imaging. J Appl Physiol (1985) 1994; 76:2076.
52. Eldridge FL. Central integration of mechanisms in exercise hyperpnea. Med Sci Sports Exerc 1994; 26:319.
53. Evans AB, Tsai LW, Oelberg DA, et al. Skeletal muscle ECF pH error signal for exercise ventilatory control. J
Appl Physiol (1985) 1998; 84:90.
54. Oelberg DA, Evans AB, Hrovat MI, et al. Skeletal muscle chemoreflex and pHi in exercise ventilatory control. J
Appl Physiol (1985) 1998; 84:676.
55. Caiozzo VJ, Davis JA, Ellis JF, et al. A comparison of gas exchange indices used to detect the anaerobic
threshold. J Appl Physiol Respir Environ Exerc Physiol 1982; 53:1184.
56. Medoff BD, Oelberg DA, Kanarek DJ, Systrom DM. Breathing reserve at the lactate threshold to differentiate a
pulmonary mechanical from cardiovascular limit to exercise. Chest 1998; 113:913.
57. Carter R, Peavler M, Zinkgraf S, et al. Predicting maximal exercise ventilation in patients with chronic obstructive
pulmonary disease. Chest 1987; 92:253.
58. Campbell SC. A comparison of the maximum voluntary ventilation with the forced expiratory volume in one

06/01/2016 20:05

Exercise physiology

14 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

second: an assessment of subject cooperation. J Occup Med 1982; 24:531.


59. Dillard TA, Hnatiuk OW, McCumber TR. Maximum voluntary ventilation. Spirometric determinants in chronic
obstructive pulmonary disease patients and normal subjects. Am Rev Respir Dis 1993; 147:870.
60. Rochester DF. Tests of respiratory muscle function. Clin Chest Med 1988; 9:249.
61. Olafsson S, Hyatt RE. Ventilatory mechanics and expiratory flow limitation during exercise in normal subjects. J
Clin Invest 1969; 48:564.
62. Kanarek DJ, Hand RW. The response of cardiac and pulmonary disease to exercise testing. Clin Chest Med
1984; 5:181.
63. Jones NL. Normal values for pulmonary gas exchange during exercise. Am Rev Respir Dis 1984; 129:S44.
64. Hansen JE, Sue DY, Wasserman K. Predicted values for clinical exercise testing. Am Rev Respir Dis 1984;
129:S49.
65. Casaburi R, Daly J, Hansen JE, Effros RM. Abrupt changes in mixed venous blood gas composition after the
onset of exercise. J Appl Physiol (1985) 1989; 67:1106.
66. Lindinger MI. Origins of [H+] changes in exercising skeletal muscle. Can J Appl Physiol 1995; 20:357.
67. Arena R, Myers J, Abella J, et al. Development of a ventilatory classification system in patients with heart failure.
Circulation 2007; 115:2410.
68. Dempsey JA, Hanson PG, Henderson KS. Exercise-induced arterial hypoxaemia in healthy human subjects at sea
level. J Physiol 1984; 355:161.
69. Hsia CC. Respiratory function of hemoglobin. N Engl J Med 1998; 338:239.
70. ERS Task Force, Palange P, Ward SA, et al. Recommendations on the use of exercise testing in clinical practice.
Eur Respir J 2007; 29:185.
71. Fishman A, Martinez F, Naunheim K, et al. A randomized trial comparing lung-volume-reduction surgery with
medical therapy for severe emphysema. N Engl J Med 2003; 348:2059.
72. Wensel R, Opitz CF, Anker SD, et al. Assessment of survival in patients with primary pulmonary hypertension:
importance of cardiopulmonary exercise testing. Circulation 2002; 106:319.
73. Kawut SM, O'Shea MK, Bartels MN, et al. Exercise testing determines survival in patients with diffuse
parenchymal lung disease evaluated for lung transplantation. Respir Med 2005; 99:1431.
74. Bassett DR Jr, Howley ET. Limiting factors for maximum oxygen uptake and determinants of endurance
performance. Med Sci Sports Exerc 2000; 32:70.
75. Koch B, Schper C, Ittermann T, et al. Reference values for cardiopulmonary exercise testing in healthy
volunteers: the SHIP study. Eur Respir J 2009; 33:389.
76. Jones NL, Makrides L, Hitchcock C, et al. Normal standards for an incremental progressive cycle ergometer test.
Am Rev Respir Dis 1985; 131:700.
77. Vogel JA, Patton JF, Mello RP, Daniels WL. An analysis of aerobic capacity in a large United States population. J
Appl Physiol (1985) 1986; 60:494.
78. Hermansen L, Saltin B. Oxygen uptake during maximal treadmill and bicycle exercise. J Appl Physiol 1969; 26:31.
79. Faulkner JA, Roberts DE, Elk RL, Conway J. Cardiovascular responses to submaximum and maximum effort
cycling and running. J Appl Physiol 1971; 30:457.
80. Woodson RD, Wills RE, Lenfant C. Effect of acute and established anemia on O2 transport at rest, submaximal
and maximal work. J Appl Physiol Respir Environ Exerc Physiol 1978; 44:36.
81. Wilson JR, Martin JL, Schwartz D, Ferraro N. Exercise intolerance in patients with chronic heart failure: role of
impaired nutritive flow to skeletal muscle. Circulation 1984; 69:1079.
82. Lewis GD, Shah R, Shahzad K, et al. Sildenafil improves exercise capacity and quality of life in patients with
systolic heart failure and secondary pulmonary hypertension. Circulation 2007; 116:1555.
83. Klausen K, Secher NH, Clausen JP, et al. Central and regional circulatory adaptations to one-leg training. J Appl
Physiol Respir Environ Exerc Physiol 1982; 52:976.
84. Casaburi R, Bhasin S, Cosentino L, et al. Effects of testosterone and resistance training in men with chronic
obstructive pulmonary disease. Am J Respir Crit Care Med 2004; 170:870.
85. Fitts RH. Cellular mechanisms of muscle fatigue. Physiol Rev 1994; 74:49.

06/01/2016 20:05

Exercise physiology

15 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

86. Hainaut K, Duchateau J. Muscle fatigue, effects of training and disuse. Muscle Nerve 1989; 12:660.
87. Systrom DM, Kanarek DJ, Kohler SJ, Kazemi H. 31P nuclear magnetic resonance spectroscopy study of the
anaerobic threshold in humans. J Appl Physiol (1985) 1990; 68:2060.
88. Metzger JM, Fitts RH. Role of intracellular pH in muscle fatigue. J Appl Physiol (1985) 1987; 62:1392.
89. Lindinger MI, Heigenhauser GJ, McKelvie RS, Jones NL. Blood ion regulation during repeated maximal exercise
and recovery in humans. Am J Physiol 1992; 262:R126.
90. Meyer RA, Dudley GA, Terjung RL. Ammonia and IMP in different skeletal muscle fibers after exercise in rats. J
Appl Physiol Respir Environ Exerc Physiol 1980; 49:1037.
91. Graham TE, Turcotte LP, Kiens B, Richter EA. Effect of endurance training on ammonia and amino acid
metabolism in humans. Med Sci Sports Exerc 1997; 29:646.
92. Lewis SF, Haller RG, Cook JD, Nunnally RL. Muscle fatigue in McArdle's disease studied by 31P-NMR: effect of
glucose infusion. J Appl Physiol (1985) 1985; 59:1991.
93. Rogers MA, Evans WJ. Changes in skeletal muscle with aging: effects of exercise training. Exerc Sport Sci Rev
1993; 21:65.
94. Chiappa GR, Roseguini BT, Alves CN, et al. Blood lactate during recovery from intense exercise: impact of
inspiratory loading. Med Sci Sports Exerc 2008; 40:111.
95. Timmons JA. Variability in training-induced skeletal muscle adaptation. J Appl Physiol (1985) 2011; 110:846.
96. Kiens B, Essen-Gustavsson B, Christensen NJ, Saltin B. Skeletal muscle substrate utilization during submaximal
exercise in man: effect of endurance training. J Physiol 1993; 469:459.
97. Rckl KS, Witczak CA, Goodyear LJ. Signaling mechanisms in skeletal muscle: acute responses and chronic
adaptations to exercise. IUBMB Life 2008; 60:145.
98. Phillips SM, Green HJ, Tarnopolsky MA, et al. Effects of training duration on substrate turnover and oxidation
during exercise. J Appl Physiol (1985) 1996; 81:2182.
99. Joyner MJ, Green DJ. Exercise protects the cardiovascular system: effects beyond traditional risk factors. J
Physiol 2009; 587:5551.
100. Gielen S, Schuler G, Adams V. Cardiovascular effects of exercise training: molecular mechanisms. Circulation
2010; 122:1221.
101. Hambrecht R, Wolf A, Gielen S, et al. Effect of exercise on coronary endothelial function in patients with coronary
artery disease. N Engl J Med 2000; 342:454.
Topic 1433 Version 6.0

06/01/2016 20:05

Exercise physiology

16 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

GRAPHICS
Muscle fiber characteristics
Type

Color

Contraction

Metabolism

Myoglobin content

Red

Slow twitch

Oxidative

High

IIA

Red

Fast twitch

Fast oxidative

High

IIX

White

Fast twitch

Anaerobic glycolysis

Low

Reproduced with permission from: Wasserman K, Hansen JE, Sue DY, et al. Physiology of exercise. In:
Principles of exercise testing and interpretation: Including pathophysiology and clinical applications, 5th ed,
Lippincott Williams & Wilkins, 2011. Copyright 2011 Lippincott Williams & Wilkins. www.lww.com.
Graphic 62260 Version 3.0

06/01/2016 20:05

Exercise physiology

17 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

Metabolic reactions in the mitochondria

The respiratory chain is composed of four multi-subunit complexes (I, II, III, and
IV) linked by the mobile electron carriers coenzyme Q and cytochrome c. The
reduced forms of nicotinamide adenine dinucleotide (NADH) and flavin adenine
dinucleotide (FADH 2 ) are formed from the citric acid cycle and the beta-oxidation
of fatty acids in the mitochondrial matrix. The respiratory chain transfers electrons
from NADH (via complex I) and from reduced flavoproteins (via complex II and
electron transfer flavoprotein-coenzyme Q oxidoreductase [ETF-Qo]) to coenzyme
Q, then complex III, cytochrome c and finally complex IV. At the same time,
complexes I, III, and IV pump electrons across the inner mitochondrial membrane
from the matrix to the intermembrane space. The influx of these electrons (protons)
back into the mitochondrial matrix releases energy that is used in the
phosphorylation of ADP to ATP by complex V (ATP synthetase), which is also
embedded in the inner mitochondrial membrane.
Graphic 67253 Version 10.0

06/01/2016 20:05

Exercise physiology

18 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

Schem reg glucose metab

Reproduced with permission from: Cryer PE, Polonsky KS. Glucose homeostasis and
hypoglycemia. In: Williams Textbook of Endocrinology, 9th ed, Wilson, JD, Foster,
DW, Kronenberg, HM, Larsen, PR (Eds), WB Saunders Co., Philadelphia 1998. p.940.
Copyright 1998 Elsevier Science.
Graphic 82039 Version 2.0

06/01/2016 20:05

Exercise physiology

19 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

Metabolic pathways of glycogen metabolism and glycolysis

The figure shows the sites of enzymatic defects resulting in clinical glycogenoses. The
glycogen storage disease types are as follows:
Type 0: Glycogen synthase deficiency
Type Ia: Glucose 6-phosphatase (G6Pase) deficiency or von Gierke disease
Type II: Acid maltase deficiency or Pompe disease
Type III: Glycogen debrancher deficiency
Type IV: Glycogen branching deficiency or Andersen disease
Type V: Muscle phosphorylase deficiency or McArdle disease
Type VI: Liver phosphorylase deficiency or Hers disease
Type VII: Phosphofructokinase (PFK) deficiency or Tarui disease
Type IX: Phosphorylase b kinase (PBK) deficiency
Type X: Phosphoglycerate mutase (PGAM2) deficiency
Type XI: Lactate dehydrogenase (LDH) deficiency
Type XII: Aldolase A deficiency
Type XIII: Beta-enolase deficiency
Type XIV: Phosphoglucomutase-1 (PGM1) deficiency

06/01/2016 20:05

Exercise physiology

20 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

Other deficiency syndromes, enzymes, and intermediates include the following:


Fructose-1,6-bisphosphatase (F-1,6-BP) deficiency
Glucokinase (GK)
Glucose transporter 2 (GLUT2) deficiency or Fanconi-Bickel syndrome
Phosphoglycerate kinase (PGK) deficiency
Phosphophoenol pyruvate carboxykinase (PEPCK) deficiency
Phosphorylase limit dextrin (PLD)
Pyruvate carboxylase (PC) deficiency
Pyruvate dehydrogenase (PDH)
Pyruvate kinase (PK) deficiency
Uridine diphosphoglucose (UDPG)
Uridine diphosphoglucose pyrophosphorylase (UDPG-P)
Adapted from: Griggs R, Mendell J, Miller R. Metabolic myopathies. In: Evaluation and
Treatment of Myopathies, Griggs R, Mendell J, Miller R (Eds), FA Davis Co., Philadelphia 1995.
p.247.
Graphic 81164 Version 10.0

06/01/2016 20:05

Exercise physiology

21 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

Calculation of AV oxygen difference and cardiac output


using the Fick method

Oxygen consumption is estimated from a nomogram based on age, sex, height,


and weight. In addition, oxygen consumption can be measured directly using
exhaled breath analysis.
Hgb: serum hemoglobin (g/dL)
SaO2: arterial oxygen saturation (percent) measured from arterial blood
SvO2: mixed venous oxygen saturation (percent) measured from the pulmonary
artery in the absence of a shunt or calculated using MvO2 = [3 SVC saturation + IVC
saturation] divided by 4 if a left to right shunt is present.
Constant values: 1.34 mL O2 per gram Hgb; 10 dL per L.
Graphic 56651 Version 3.0

06/01/2016 20:05

Exercise physiology

22 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

Effect of exercise training on oxygen uptake at peak exercise

The left panel illustrates representative normal values for oxygen uptake at rest, at peak exercise, and at peak
exercise following exercise training in a 25-year-old male. The right panel illustrates relative changes in the three
components of VO 2 (heart rate, stroke volume, and arteriovenous difference in oxygen content) in response to
maximal exercise.
VO 2 : oxygen uptake; HR: heart rate; SV: stroke volume; C: content.
Courtesy of Gregory D Lewis, MD.
Graphic 95312 Version 1.0

06/01/2016 20:05

Exercise physiology

23 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

Heart rate and oxygen consumption during


exercise

Schematic representation of the heart rate - oxygen uptake (VO 2 )


relationship during incremental exercise in normal (red circles) and
well-trained individuals (blue squares).
Graphic 77997 Version 2.0

06/01/2016 20:05

Exercise physiology

24 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

Blood flow distribution at rest and at maximal exercise


Rest (percent total)

Maximal exercise (percent total)

Splanchnic

24

Skeletal muscle

21

88

Kidneys

19

Brain

13

Skin

Heart

Other organs

10

Graphic 62660 Version 1.0

06/01/2016 20:05

Exercise physiology

25 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

Oxyhemoglobin dissociation curve

Depicted here is the oxyhemoglobin dissociation curve for normal adult


hemoglobin (Hemoglobin A, solid line). Note that hemoglobin is 50 percent
saturated with oxygen at a partial pressure of 27 mmHg (ie, the P50 is 27
mmHg) and is 100 percent saturated at a PaO 2 of approximately 100 mmHg.
Depicted here are curves that are "left-shifted" (blue line, representing
increased oxygen affinity) and "right-shifted" (red line, decreased oxygen
affinity). The effect of right- or left-shifting of the curve is most pronounced at
low oxygen partial pressures. In the examples shown, the left-shifted curve
means that hemoglobin can deliver approximately 70 percent of its attached
oxygen at a PaO 2 of 27 mmHg. In contrast, the right-shifted hemoglobin can
deliver only about 35 percent of its attached oxygen at this PaO 2 . A high
proportion of fetal hemoglobin, which has high oxygen affinity, shifts this
curve to the left in newborns.
Graphic 81216 Version 5.0

06/01/2016 20:05

Exercise physiology

26 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

Ventilatory threshold

The ventilatory threshold is the intensity at which ventilation begins to


rise in excess of VO 2 . It is best demonstrated by plotting VE/VO 2 and
VE/VCO 2 on the same graph (as above). The ventilatory threshold is the
point just before VE/VO 2 begins to rise without a similar rise in
VE/VCO 2 .
VE: volume of gas expired (L/min); VO 2 : volume of oxygen consumed
(L/min); VCO 2 : volume of carbon dioxide produced (L/min).
Graphic 52387 Version 3.0

06/01/2016 20:05

Exercise physiology

27 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

Ventilatory efficiency predicts risk for major cardiac events

Kaplan-Meier analysis for two-year major cardiac-related events. Subjects meeting criteria for
VC-1 (VE/VCO 2 slope 29.9; n = 144) experienced four major cardiac events (including two
heart transplants); 97.2% were event-free. Subjects meeting VC-II criteria (VE/VCO 2 slope 30 to
35.9; n = 149) experienced 22 major cardiac events (including three LVAD implantations);
85.2% were event-free. Subjects meeting VC-III criteria (VE/VCO 2 slope 36 to 44.9; n = 112)
experienced 31 major cardiac events (including three LVAD implantations and two heart
transplants); 72.3% were event-free. Subjects who met VC-IV criteria (VE/VCO 2 slope 45; n =
43) experienced 24 major cardiac events (including two LAVD implantations and five heart
transplants); 44.2% were event-free. Log-rank 86.8, p<0.0001.
VE: minute ventilation; VCO 2 : carbon dioxide output; LVAD: left ventricular assist device.
From: Arena R, Myers J, Abella J, et al. Development of a ventilatory classification system in patients
with heart failure. Circulation 2007; 115:2410. Reproduced with permission from Lippincott Williams &
Wilkins. Copyright 2006 American Heart Society. Unauthorized reproduction of this material is
prohibited.
Graphic 95314 Version 1.0

06/01/2016 20:05

Exercise physiology

28 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

Oxygen consumption and workload during exercise

Schematic representation of the normal VO2-workload relationship


during incremental exercise to exhaustion (blue circles). The
relationship is left-shifted for obese individuals (red triangles).
Graphic 64164 Version 1.0

06/01/2016 20:05

Exercise physiology

29 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

Normal values for maximum oxygen uptake during cycle ergometry


Age

25-34

35-44

45-54

55-64

>64

Men
Median VO 2
max
(mL/minute
per kg)
(5 th-95 th
percentile)

38
(26-51)

36
(24-48)

35
(23-45)

33
(20-40)

30
(17-36)

30
(22-41)

29
(21-39)

27
(20-37)

25
(18-32)

22
(16-28)

Women
Median VO 2
max
(mL/minute
per kg)
(5 th-95 th
percentile)

Normal VO 2 max values, based on incremental bicycle ergometry testing in healthy individuals
derived from a population-based cohort.
VO 2 max: maximum oxygen uptake during exercise.
Data from: Koch B, Schaper C, Ittermann T, et al. Reference values for cardiopulmonary exercise testing in
healthy volunteers: the SHIP study. Eur Respir J 2009; 33:389.
Graphic 95298 Version 1.0

06/01/2016 20:05

Exercise physiology

30 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

Lactate threshold during exercise

Schematic representation of the normal blood lactate - oxygen uptake


(VO 2 ) relationship during incremental exercise. The lactate threshold
(LT) denotes the onset of a sustained rise in blood lactate.
Graphic 78425 Version 2.0

06/01/2016 20:05

Exercise physiology

31 sur 31

http://www.uptodate.com/contents/exercise-physiology?topicKey=PULM...

Disclosures
Disclosures: David M Systrom, MD, FRCPC Nothing to disclose. Gregory D Lewis, MD Nothing to disclose. James K Stoller, MD, MS
Consultant/Advisory Boards: CSL Behring [Alpha-1 antitrypsin deficiency (Augmentation therapy (IV alpha-1 antiprotease))]; Kamada [Alpha-1
antitrypsin deficiency (Augmentation therapy (IV alpha-1 antiprotease))]; Baxter [Alpha-1 antitrypsin deficiency (Augmentation therapy (IV
alpha-1 antiprotease))]; Grifols [Alpha-1 antitrypsin deficiency (Augmentation therapy (IV alpha-1 antiprotease))]; Boehringer Ingelheim [COPD
(Bronchodilators (various))]; Arrowhead Research [Alpha-1 antitrypsin deficiency]. Helen Hollingsworth, MD Nothing to disclose.
Contributor disclosures are reviewed for conflicts of interest by the editorial group. When found, these are addressed by vetting through a
multi-level review process, and through requirements for references to be provided to support the content. Appropriately referenced content is
required of all authors and must conform to UpToDate standards of evidence.
Conflict of interest policy

06/01/2016 20:05

S-ar putea să vă placă și