Chapter 14: Molecular Basis of Exercise Training Adaptations in Skeletal Muscle

David Hood; Nashwa Cheema; Ayesha Saleem; Heather Carter; Anna Vainshtein; Olga Ostojic; and Sobia Iqbal

Photo 14-1.  Strong, wellness couple doing kettlebell weight exercise, workout or training inside a gym. Resistance exercise activates specific genes in our muscles. Photo by peopleimages.

Learning Objectives

  • Understand the concept of adaptation as it relates to changes in gene expression;
  • Appreciate the divergent signals and pathways that lead to adaptation in exercising muscle;
  • Comprehend the details of gene expression and be able to converse about how changes in muscle adaptations apply to a real training setting.

 

Key Terms

muscle adaptation, gene expression, transcription, translation, protein synthesis, myosin heavy chain, mitochondrial biogenesis biogenesis, fibre types, electron transport chain, isoforms, NuGEMPs, transgenic, hypertrophy, endurance training, resistance training training, proteins, satellite cell, myonuclear domain, slow twitch, fast twitch, receptors (DHPR), creatine kinase (CK), myoglobin, heat-shock protein (HSP)-70, ROS, PGC-1α, Nuclear Respiratory Factor-1 (NRF-1), mTOR, nuclear factor of activated T-cells (NFAT), S6Kinase, IGF, SERCA, O2max, ADP.

 

Case Presentation: A Look Inside the Muscle: The Metabolic Benefits of Training

Andrew is a 26-year-old university graduate who has never been an elite athlete. In fact, he was recreationally active in only a few sports during high school and university, preferring to focus on his studies and academic achievements. His girlfriend Erin is sports-minded, and always participated in track and field in her youth. Now she finds running the occasional 10 km race rewarding since her track career has ended. Erin has convinced Andrew that his fitness level is on the low end of average, and that starting a training regimen (jogging and running) with the goal of completing a running race together might be a way to spend time together as well as to get fit. In fact, through her university connections, she has arranged a complete “work-up” and analysis of Andrew’s fitness, from the level of the muscle to his aerobic capacity. During a pre-test,O2max would be assessed and micro-biopsy samples would also be taken to analyze the metabolic changes within the muscle during exercise before and after the training period. Erin has set up a progressive training program appropriate for Andrew’s starting fitness level beginning about 10 weeks prior to the race. Andrew underwent a graded exercise stress test before and after the training period in which and muscle biopsy samples are taken at various workloads.

Photo 14-2 A group of researchers assess the aerobic fitness of a male subject like Andrew. By Fretamales [CC BY-SA 3.0 (https://creativecommons.org/licenses/by-sa/3.0) or GFDL (http://www.gnu.org/copyleft/fdl.html)], from Wikimedia Commons

The results of Andrew’s training are illustrated in Figure 14-1. They are typical of changes from a previously untrained state to that of a more physically fit individual who faithfully adheres to a well-designed endurance training program. Andrew’s O2max increased by about 10% and the proportion of his type IIa fibres increased at the expense of type IIx fibres. His proportion of type I fibres did not change, as expected. Typically, a transition to more type I fibres would take a much longer period of training (see later in this chapter). During graded exercise, Andrew displayed the beneficial metabolic adaptations of training, corresponding to improved endurance performance. Andrew used less glycogen during exercise and produced less lactate during the submaximal portion of the stress test. He also spared his use of phosphocreatine and the level of adenosine diphosphate (ADP) was reduced. This lower ADP concentration was a reason for the lower level of glycolytic stimulation. Mitochondrial marker succinate dehydrogenase was elevated, indicating that the training program increased mitochondrial content within the trained muscle. This adaptation would account for a better use of lipid as a substrate during exercise and contribute to the preservation of carbohydrate (COH) fuels like glycogen which must be preserved in order to perform well in endurance events. Andrew’s muscles have a higher oxidative capacity because they have more mitochondria. They also have more capillaries surrounding each muscle fibre which provide energy substrates and oxygen to the mitochondria. His endurance time is much more prolonged, and he is ready to run his race with Erin!

This figure presents a series of graphs; four are histograms and 4 are scatter plots with lines joining the lines. The top left histogram shows the change in fibre type as a result of the training. Type IIa fibres have increased while Type IIx have decreased. Maximal oxygen uptake increased slightly from 40 to 45 ml/kg/min.Biopsy samples showed a small increase in succinate dehydrogenase activity and the ratio of capillaries to muscle fibres increased from 2.5 to 4.4. During the incremental test, Andrew was able to go an extra stage without an increase in blood lactate. Muscle glycogen use during the incremental test was reduced, and the decrease in creatine phosphate concentration was considerably less. Similarly, the rise in ADP concentration was less after the training program.
Figure 14-1 (Top) Changes in muscle fibre type,V̇O2max, muscle succinate dehydrogenase (SDH) activity and capillary: fibre ratio pre- and post-training. Line graphs illustrate the changes in metabolites (lactate, glycogen, phosphocreatine (PCr) and ADP) in muscle during graded exercise up to V̇O2max both pre- and post-training.

 

Gender BOX: Battle of the Sexes? It’s a Molecular Matter!

There is no denying that males and females differ vastly in their muscle structure, size, and ability to undergo hypertrophy in response to training. Males feature a greater free mass, a bigger muscle cross-sectional area, and more fast-twitch type IIA muscle fibres when compared with their female counterparts (Miller et al. 1993). Females, on the other hand, display superior resistance (Wüst et al. 2008) and a greater reliance on for energy production (Roepstorff et al. 2006). It has been postulated that these discrepancies are due to inherent hormonal differences between the two genders where men have 10-30-fold greater resting levels of testosterone while women are well-endowed with estrogen. Testosterone has the ability to enhance protein synthesis, ultimately contributing to muscle hypertrophy, while estrogen activates the synthesis of burning machinery. Does this mean that men adapt better to exercise training than women? Not necessarily! Several studies indicate that men and women undergo the same level of protein synthesis following an acute bout of resistance exercise despite dramatic differences in testosterone levels (West et al. 2012). Men and women also adapt similarly to high intensity interval training (Astorino et al. 2011). However, if the two sexes display similar adaptations to resistance training what accounts for the extensive increase in muscle size observed in men? Other than the hormonal differences, there also appears to be variations in gene expression between the two genders. The muscle transcriptome is vastly different between the two sexes both at baseline and following an acute bout of resistance exercise (Liu et al. 2010). The time course of the transcriptional modulation is also sex-dependent wherein males experience prolonged changes in gene expression while females exhibit rapid transitions (Liu et al. 2010). Thus, in addition to the hormonal divergence, transcriptional regulation might be at the root of the differential adaptations observed between the sexes following exercise training at the same relative intensity.

 

Introduction to the Types of Exercise Training

Exercise can be classified in a number of ways. Acute exercise refers to a single bout which can either be short (i.e. seconds) or long (several hours), and can vary in intensity from being mild, as in a light jog, to supramaximal, like a 100-meter sprint at top speed. A single acute exercise bout produces physiological and metabolic changes in the working muscle that, if repeated many times, can ultimately lead to an “adaptation”. When acute exercise bouts are repeated over and over again, this is referred to as exercise training. In order for exercise training to be effective, certain training principles must be adhered to. These include:

  • the frequency of exercise per week;
  • the duration of each exercise bout in minutes;
  • the intensity of exercise, usually expressed in terms of %O2max, or %max heart rate;
  • the length of the training period in weeks or months.

These training principles should be applied regardless of the type of training employed. Traditionally, three forms of training have been recognized for many years. As reviewed below, these are:

Endurance Training

This type of training usually refers to repeated bouts of acute endurance exercise which employs a large fraction of available muscle mass. Commonly accepted exercise dosage for adults aged 18-64 years is to engage in moderate-to-vigorous intensity aerobic exercise (65-75% O2max) for 30-40 minutes at a frequency of 3-5 days per week. More recently, the Canadian Society for Exercise Physiology has revised these guidelines to indicate that adults should aim to accumulate 150 minutes of exercise per week in bouts of 10 minutes or more. From a health perspective, this helps reduce the risk of chronic pathological maladies such as heart disease, stroke, obesity, type 2 diabetes and osteoporosis. Typical examples include brisk walking, jogging, running, swimming, cycling and cross-country skiing. These types of exercise produce transient increases in heart rate, cardiac output and peripheral muscle oxygen and substrate extraction, leading to an acute increase in above rest to a steady-state level which can be sustained for a prolonged period. When this is repeated 3-5 times per week for several weeks, endurance training benefits such as improved whole-body endurance and an increase in max, will accrue. The increase in endurance is a result of muscle adaptations to exercise, as outlined in this chapter.

Resistance Training

Photo 14-3 Resistance exercise involves resisting body movements, often using weights. Generally, this type of exercise is repeated 3-4 times per week, with a rest day in between sessions. Exercise is performed using specific muscle groups, such as the biceps, pectoral muscles, or quadriceps, for example. Typically, 3 “sets” of exercise are performed, with each set consisting of 8-20 repetitions of the same motion (e.g. biceps curl) with a given weight. The lower number of repetitions (i.e. 8) generally employs a higher load in order to maximize strength gains, while the higher number of repetitions uses a smaller load and leads to greater improvements in muscular endurance. The load chosen is usually expressed as a percentage of the 10 Repetition Maximum (10RM), or the maximum load that can be moved in ten repetitions of the exercise. Resistance exercise performed over the course of several weeks to months leads to an increase in size and strength of the specific muscle groups that are trained. Improvements in muscle mass are directly related to improving quality of life, especially in the elderly population where muscle frailty often leads to falls and a subsequent dependence on others to perform simple daily tasks. The increase in muscle mass is due to adaptations brought about by the high forces generated within the fibres recruited during the resistance training. These adaptations are different from those observed with endurance training, as described below.

Interval Training

Photo 14-4 Interval training involves repetitive bouts of intense aerobic exercise followed by periods of rest.

This type of training has been well recognized for many years, and it has regained some popularity recently because of the adaptations that can be observed within muscle. Interval training usually entails the performance of vigorous or supramaximal intensity exercise over short intervals of time. An example of such training is repeated 400-meter sprints at a pace of 60 seconds (6.7 metres·s-1), done with a rest interval of 2 minutes between sprints. This represents a “work-to-rest” ratio of 1:2. Various work-to-rest ratios can be employed depending on the fitness of the subjects. Many, many different types of intervals can be imagined, tailored to the specificity of the sport or activity. A longer rest interval may permit a higher intensity in the next interval. During high intensity interval training, it is more likely that the rarely used Type IIx motor units within the muscle will be recruited. The fibres within these motor units adapt readily to exercise because they have a very low endurance to begin with. Not only are muscles and motor units within them heavily activated by such high intensity exercise but so are the respiratory and cardiovascular systems. High heart rates are common, and if the rest interval is short enough, the heart rate remains high even during the recovery period between bouts. This maximizes the cardiovascular benefits of such exercise. Known for its difficulty, interval training has recently seen a revival of interest from exercise physiologists who wish to minimize time spent training and maximize the benefits of high intensity exercise.

Subcellular Systems that Adapt to Training

Like most cell types in the body, muscle cells contain the organelles responsible for cell survival including a nucleus, ribosomes, endo- and sarcoplasmic reticulum, lysosomes, cytoskeletal elements and a cytoplasm with an abundance of proteins. The specialized elements within muscle cells that distinguish them from many other cell types include

  • multiple myonuclei;
  • a well-developed calcium handling system that responds with every action potential;
  • myofibrillar proteins that produce force;
  • three energy systems that provide rapid and sustained for muscle contractions.

Myonuclei

During embryonic development, stem cells commit themselves to form muscle cell precursors called myoblasts. Myoblasts have a single nucleus and lack proteins. Myoblasts then fuse together to form myotubes which are elongated immature muscle cells. When many myoblasts fuse together, they donate their individual nuclei to the growing, elongated myotube, forming a multinucleated muscle cell (Figure 14-2). As progresses, these nuclei express muscle-specific

Two pictures of histology slides are shown. The top one is of several small myoblasts. Each cell has an outer membrane and a large black nucleus. Cytoplasm is stained green. The distribution of myoblasts appears to be random. In the lower picture, the myoblasts have aligned and fused, forming immature muscle cells or myotubes. Each myotube has several centralized nuclei.
Figure 14-2 Image of myoblasts (Top) and differentiated immature muscle cells (myotubes) grown in cell culture. The dark circles represent individual myonuclei within the cells. Myoblasts have fused together to produce multi-nucleated myotubes in the lower figure.

genes, such as and myosin, and the cell gains the ability to contract. Nuclei are necessary all along the muscle fibre as a single nucleus is only able to provide the DNA and building blocks necessary to sustain a certain volume of the cell. This volume is known as the myonuclear domain and it corresponds to the volume of cytoplasm governed by each individual nucleus (Figure 14-3). Multiple nuclei are necessary to provide sufficient machinery to support the full length of the muscle fibre. Thus, unlike most cells in the body which possess only one nucleus, mature muscle fibres have many. As we discuss muscle adaptations to exercise, we will continually refer to the genes within these nuclei and how these genes are “expressed” to give rise to specialized adaptations that we observe with training. In addition, as discussed below, the number of myonuclei in muscle can be altered by training, giving the cell the potential to increase gene expression to better support the needs of a growing muscle.

A segment of a muscle fibre is drawn in the top of this image and two histology sections of fibre segments are shown below. The diagram has blue nuclei on the surface with dashed lines drawn in a way to separate each nucleus, indicating an area or domain that belongs to each nucleus. At the top and bottom are two quiescent satellite cells with blue nucleus and green cytoplasm. The basement membrane encloses these cells, but they are clearly outside the muscle membrane. The two fibres shown below present with very different volume per nucleus. On the left, there are many more nuclei than on the right, for similar sized fibre segments. An arrow points to a yellow nucleus on the right fibre, identifying this as a nucleus that is apoptotic (undergoing degradation).
Figure 14-3 (Top) Illustration of a segment of a muscle fibre with its theoretical myonuclear domains, along with adjacent quiescent satellite cells. (Bottom) Photograph of fibre segments with small (left) and large (right) myonuclear domains; a rare apoptotic myonucleus is stained yellow (arrow). Adapted from Saleem A. et al., in Reed J. and D. Green, Eds. Apoptosis: Physiology and Pathology. New York: Cambridge, 2011, p. 318.

Myofibrillar proteins

Within the cytoplasm of muscle cells are organized bundles of proteins called myofibrils . These myofibrils contain

  • proteins,
  • structural proteins,
  • cytoplasmic proteins,
  • regulatory proteins.

Contractile proteins generally refer to and myosin, while structural proteins include titin, z-line proteins and nebulin. Regulatory proteins are those which “control” the contraction process. These include troponin and tropomyosin (Figure 14.4). Both the contractile and regulatory proteins exist as multiple “isoforms” within muscle fibres. Isoforms are variants of the same protein, but with a slightly different primary amino acid sequence. This gives them a different physiological capability within the cell. Some physiologically-relevant isoforms that exist within muscle are listed in Table 14-1.

Mammals possess four genes that encode MHC isoforms that are expressed in adulthood. These are the MHC Type I, IIa, IIx and IIb genes (Figure 14-5). The MHC isoforms were discovered based on their different molecular weights when they are separated using a technique called electrophoresis, which distinguishes proteins based on their size. It is now known that only three of these four genes are expressed in adult human muscle and that the fourth, the MHC Type IIb, is only expressed in smaller mammals, like rabbits, rats and mice. The abundance of these proteins in muscle fibres can be very high, giving rise to the idea that muscle “fibre types” can be classified based on how much of each isoform is present. For example, a  type IIa fibre contains mostly MHC IIa. Because MHC IIa is a fast isoform of MHC, this means that a type IIa fibre is a fast-twitch fibre. In contrast, a fibre composed of mostly MHC I is a slow-twitch fibre because this MHC has a slow rate of breakdown. The abundance of these isoforms is typically detected using histochemistry which yields a qualitative analysis of the muscle fibres (Figures 14-5, 14-6). However, more quantitative analyses based on single fibre electrophoresis reveals that many muscle fibres, referred to as hybrid fibres, express multiple MHC isoforms. Our classification of Type I, IIa and IIx fibres is really an oversimplification of what really exists in muscle. Nonetheless, the classification into these three fibre types remains useful in general.

A question that often arises in exercise physiology is “Can fibre types be changed with training”? To do this, one would have to train at a sufficient exercise intensity, duration and frequency, and for a long period of time to induce a change in the expression of the MHC genes. Once we learn about the gene expression pathway (below), this fascinating topic will be considered once again.

Sarcoplasmic Reticulum (SR) Proteins

As discussed in earlier chapters on muscle contraction, when an action potential within a muscle fibre occurs, the change in voltage across the plasma membrane propagates along the plasmalemma and down the t-tubule which is continuous with the plasmalemma (Figure 14-4). Embedded in the t-tubule membrane are proteins called dihydropyridine receptors (DHPR). These are voltage sensitive and alter their conformation when an action potential occurs. The DHPRs physically interact with the calcium release in the SR, also called the ryanodine receptors (RyR). An action potential causes conformational changes in these two proteins including the opening of the RyR to permit a burst of calcium to be released into the cytoplasm. When action potentials cease, calcium release channels close and the calcium in the cytoplasm is taken back up into the SR in an ATP-dependent process catalyzed by the sarco/endoplasmic reticulum calcium- ATPase (SERCA) which, like MHCs, also exist in different isoforms. SERCA1a is the primary isoform present in adult fast-twitch skeletal muscle whereas SERCA2a is highly expressed in slow-twitch muscle and heart. SERCA3 exists in many cells of the body. SERCA-mediated calcium uptake allows cytoplasmic calcium to return to resting levels and causes the muscle to relax (Figure 14-4).

A number of these proteins have the potential to adapt to training though the study of SR proteins is not as advanced as for the myofibrillar or metabolic proteins. Endurance exercise training appears to reduce SR calcium uptake and release. Several proteins involved in SR calcium cycling undergo alterations resulting in this effect. For example, decreases in the content of the fast isoform SERCA1a and the RyR protein have been noted, which would tend to reduce the speed at which calcium is handled within endurance-trained muscle.

The sequence of excitation-contraction coupling is illustrated in this cartoon. At the top of the image, the extracellular space is shown, with a membrane separating the sarcoplasm from this extracellular space. A transverse tubule is shown, protruding down into the cell. An action potential is shown at the membrane with an arrow indicating propagation along the surface. Within the transverse tubule, a dihydropiridine receptor is embedded in the membrane. This receptor is linked to the ryanodine receptor in the membrane of the sarcoplasmic reticulum. Arrival of the action potential is sensed by the dihydropyridine receptor, which opens the ryanodine receptor allowing diffusion of Ca2+ from the sarcoplasmic reticulum into the cytoplasm. The myofilaments are drawn within the muscle fibre and an expanded view of these shows Ca2+ binding to troponin, which is connected to the action of the thin filament. This triggers contraction, and a twitch occurs. The twitch is illustrated as a small graph on the left. Force rises and falls over a period of milliseconds. A blue twitch represents a fast-twitch fibre, rising and falling very quickly. A red twitch represents a slow-twitch fibre, rising more slowly then falling. A transporter known as SERCA transports Ca2+ back into the sarcoplasmic reticulum, restoring the relaxed condition.
Figure 14-4 Review of the contractile response to an action potential. Voltage sensor proteins such as the DHPR change conformation in response to an action potential and cause the calcium release (Ryanodine receptor; RyR) to open, allowing calcium to escape the sarcoplasmic reticulum (SR). Calcium diffuses to the myofilaments where it interacts with the Troponin-C, allowing tropomyosin to remove its inhibition of the myosin-binding site on actin, allowing for actin-myosin interaction, breakdown, and crossbridge cycling. When the action potential terminates, calcium is no longer released and cytoplasmic calcium can diffuse back the SR, aided by the transport protein parvalbumin (PV), thus promoting muscle fibre relaxation. The sarco-, endoplasmic reticulum calcium ATPase pump facilitates calcium uptake into the SR in an ATP-dependent manner.

Protein

Isoforms

Functions

Distinctions

Myosin heavy chain (MHC)

MHC I

MHC IIa

MHC IIx

Catalyze the breakdown of ADP during the crossbridge cycle, when myosin is bound to actin

MHC I has a slow rate of breakdown, whereas isoforms IIa and IIx catalyze fast breakdown

Actin

Alpha-actin

Gamma-actin

α- is a protein involved in the crossbridge cycle, whereas γ- is a cytoskeletal protein involved in maintaining the shape of the cell

Lactate dehydrogenase

M-LDH

H-LDH

M-isoform catalyzes the reaction of Pyruvic acid + NADH Lactic acid + NAD+

The H-isoform catalyzes the reaction in the opposite direction

Sarcoplasmic-endoplasmic reticulum ATPase

SERCA 1

SERCA2

Catalyze the uptake of calcium back into the sarcoplasmic reticulum following the contraction phase. This uptake produces muscle relaxation

SERCA1 is found in fast-twitch, Type IIa and IIX fibers. SERCA2 is found in slow-twitch f ibers

 

This drawing illustrates the determination of fibre type in a muscle cell. In the top left is shown a strand of DNA, with coding for 4 types of myosin: type I, type IIa, type IIx and type IIb. This DNA would be present in all nucleii of the myocyte. Beneath this is a representative myocyte with multiple nucleii. Evaluation of the cell for fibre type can be by histochemistry which relies on staining a microscope slide containing the muscle sample or by electrophoresis, which requires preparation of a muscle sample and application of the sample to a channel of gel.
Figure 14-5 Multinucleated single fibers have genes encoding all four MHC isoforms. Only the type I, Iia and Iix isoforms are expressed in humans. The gene that is expressed the most will lead to the greatest abundance of MHC isoform within the fiber, and this can be determined by histochemistry staining. Alternatively, single fiber electrophoresis can provide a more quantitative description of the relative expression of each gene, sometimes revealing the co-expression of an alternative isoform within the same fibre.
A series of histochemical slides are shown for samples taken sequentially from a muscle biopsy. A illustrates the general structure of the sample: membrane bound cells with thicker bands of connective tissue streaking across the sample. The same bands of connective tissue are evident in each of the tissue slices. B shows dark cell borders where dystrophin is stained. C: of the hundreds of cells in the image, only 6 of them stain darkly, indicating very few type I fibres. D Fast-twitch fibres are stained darkly while the slow-twitch fibres that were dark in the previous slide. E shows staining for myosin ATPase type IIa. Most of the fibres are dark, but the slow fibres identified in C are not, and it is clear that there are two levels of staining in the rest of the fibres.
Figure 14-6 Serial cross-sections of rabbit tibialis anterior muscle under various staining conditions. The identical fiber can be traced through each section to compare the properties of fibers. (A) Hematoxylin and eosin, demonstrating general muscle fiber morphology. (B) Dystrophin immunohistochemistry showing the subsarcolemmal nature of this protein (brown deposits represent protein). (C) Myofibrillar ATPase under acid preincubation conditions. Under these conditions, slow fibers stain darkly as do the extracellular capillaries while fast fibers stain lightly. (D) Myofibrillar ATPase under alkaline preincubation conditions. Under these conditions, slow fibers stain lightly while fast fibers stain darkly. Note that in both panels C and D, fast fiber staining intensity occurs at two levels. (E) Immunohistochemical reaction for fast myosin heavy chain antibody. In rat skeletal muscle, this stains type 2A fibers darkly and type 2X fibers more lightly and is negative for types 2B and 1 fibers. (F) Succinate dehydrogenase (SDH) used to demonstrate muscle fiber mitochondrial content. Note that the slow fibers (Samples fiber labelled with a “1”) as well as the type 2A fast fibers (sample fiber labelled with a “2A”) have higher oxidative capacity compared to the type 2X fibers (sample fiber labelled with a “2X”). Adapted from Lieber, RL. Skeletal Muscle Structure, Function and Plasticity, 3Ed., Philadelphia: LWW, 2010.


Cytosolic Proteins

There are a number of different types of proteins within the cytoplasm but the most studied are those which are involved in energy metabolism, particularly enzymes of glycolysis or the metabolism of phosphocreatine (i.e. creatine phosphokinase, or CPK). These enzymes have a very high specific activity in muscle fibres, much higher (10-fold!) than those enzymes devoted to oxidative metabolism in the mitochondria, for example. A large number of studies have been conducted which have evaluated the effects of training on the activity of these enzymes and most indicate very little effect, regardless of whether endurance, sprint or resistance training methods are employed.

In addition to metabolic enzymes, two other proteins which serve a “chaperone”-like role are responsive for changes in muscle activity. These are parvalbumin, a calcium binding protein, and myoglobin (Mb), a protein which binds oxygen. These proteins are chaperones in the sense that they facilitate the transport of their ligands (i.e. calcium or oxygen respectively) to the site of where they are needed. For example, parvalbumin facilitates the diffusion of calcium to the SERCA pump for removal from the cytoplasm. This protein is considered to be important for the relaxation phase of muscle contraction and it is very high in concentration in fast-twitch muscle fibres and low in slow-twitch muscle fibres. Myoglobin, on the other hand, helps to transport oxygen from the periphery of the muscle fibre to the where it can be used in phosphorylation, the aerobic production of ATP. Myoglobin is highly abundant in slow-twitch and Type IIa fibres. These fibre-types have a red colour due to the presence of Mb which contains the red pigment, heme. In response to chronic endurance training, parvalbumin tends to decrease in the cell while Mb is usually found to increase (Figure 14-7).

A third class of proteins are referred to as stress proteins, typified by the heat-shock protein-70 (HSP-70). These proteins are induced very rapidly in response to a variety of stressors including exercise. Their main function appears to be in protein refolding. Any protein that is denatured, or unfolded, during a stressful cellular condition loses its function. This undermines the health of the cell. HSP70 has an active role in preventing cellular death by helping to restore protein shape and function.

Mitochondrial Proteins

Mitochondria are the “powerhouses” of the cell because they are the source of most of the that is used to fuel muscular contraction. The proteins within are those involved in the oxidation of pyruvic acid, derived from glycolysis, and β oxidation, the breakdown of fatty acids from triglycerides derived from within the muscle cell itself or originally found within adipose tissue. The proteins ultimately responsible for this metabolism are the enzymes of Krebs’ Cycle and the electron transport chain (ETC) as described earlier (Chapter 6). One of the most dramatic adaptations to regularly performed exercise training is the uniform increase of many enzymes involved in β-oxidation, Krebs’ Cycle and the ETC. This occurs in the muscle fibres that are recruited, or used, during the training period. For example, in response to training of relatively mild intensity (i.e. jogging), the muscle fibres that are predominantly used during this exercise will adapt and produce more mitochondrial proteins. These fibres would include type I and IIa, but probably not the Type IIx fibres which are only activated during intense exercise bouts. It is also important to realize that mitochondrial adaptations are lost rather rapidly once a training regime ceases (Figure 14-8). Since mitochondrial adaptations closely correlate with endurance performance and measures of aerobic fitness, it is important to maintain a level of physical activity to avoid losing these adaptive changes.

A pair of histograms are shown here, one depicting parvalbumin where in the trained state, parvalbumin is reduced from about 128 to about 78. In the second graph, Myoglobin is shown to increase from 1.2 to about 2.2 in trained muscle.
Figure 14-7 Typical endurance training-induced changes in the cytsolic proteins, parvalbumin and myoglobin in muscle.
This graph shows the percent of pretraining value for mitochondrial concentration and maximal oxygen uptake for progression through 8 weeks of training followed by 8 weeks of detraining. mitochondrial volume increases to about 140% of pretraining level after 8 weeks of training, then decreases to below the baseline value with 8 weeks of detraining. In contrast, maximal oxygen uptake increases to about 120% of pretraining value after 8 weeks of training, but only decreases to about 112 % of pretraining value after 8 weeks of detraining.
Figure 14-8 Typical changes in V̇O2max and muscle mitochondrial enzymes (MITO) induced by 8 weeks of endurance training and followed by de-training for 8 weeks. Modified from Henriksson, J. and Reitman J.S. Acta Physiol. Scand. 99: 91-97, 1977.

There are many benefits of mitochondrial adaptation including an improvement of lipid metabolism and a reduction in the reliance on COH as energy sources during prolonged exercise. This is important because of the limited (gram) quantities of COH stores within our muscles and liver. In contrast, an increase in allows us to take greater advantage of the (kilogram) quantities of stores within our adipocytes and muscle cells. In addition, having more means that we can rely less on glycolysis and phosphocreatine as energy sources during intense contraction conditions. This reduced reliance on these pathways means that we produce less lactic acid during exercise, glycogen is used at a slower rate, and phosphocreatine breakdown is diminished. These adaptations have important benefits for endurance performance. A perfect example of this is shown in the Case Study at the start of this Chapter. Traditionally, this kind of adaptation has been observed with endurance training, but more recently it has been shown to occur with different forms of sprint training, which produce adaptations in type IIx fibres as well.In order to understand how these adaptations in muscle occur, it is important the review the pathway of gene expression and how it is activated in response to different exercise stimuli. Gene expression is the precursor to protein synthesis; therefore, both processes are crucial for muscle adaption with exercise.

Highlight BOX: The “Secret” of Muscle

Within the last 10-15 years, numerous studies have shown that muscle tissue is capable of producing and releasing hundreds of different proteins (Bortoluzzi, (2006). The ability of skeletal muscle to secrete factors provides a method for communication. The signaling molecules that are released have been called “myokines” (Pedersen, 2008). These secreted signals can travel in the blood and affect other tissue types, well known as an endocrine effect. Secreted myokines may also remain local and work to cause changes within the muscle itself. This is known as an autocrine or paracrine effect. While not all of the proteins secreted from muscle have been identified yet, the IGF (insulin-like growth factor) family, IL-6 (interleukin-6), and most recently irisin (named after the Greek messenger goddess, Iris), have all been accepted as myokines.

The IGF family consists of many proteins. In skeletal muscle, IGF-I and IGF-II can be secreted and act back on the muscle in an autocrine/paracrine manner. IGF-II seems to promote the formation of muscle during development (Florini , 1991). In adults, IGF-I is capable of increasing muscle size (i.e. hypertrophy) as well as promoting muscle repair through the activation of satellite cells (Barton, 2006).

IL-6 was among one of the first- ever myokines to be identified. It was discovered that IL-6 was released from muscle during exercise, and that plasma blood levels could increase up to 100-fold! (Steensberg, 2002). This increase in IL-6 depends on both the exercise intensity and duration. Increased muscle-derived IL-6 triggers breakdown (lipolysis) and utilization (fat oxidation) in other tissues. Mice that lack IL-6 become obese and resistant (Pedersen, 2008 ).

The latest myokine to be identified is irisin. Irisin production is regulated by a protein called PGC-1α which is a critical factor for exercise adaptations in skeletal muscle (Boström , 2012). It is very well known that PGC-1α and circulating levels of irisin both increase with exercise. Once released from the muscle, irisin is able to travel to the adipose tissue where it will stimulate the to increase energy expenditure which will be released as heat. Administration of irisin to rodents is also shown to stimulate weight loss and prevent insulin resistance. Thus, production of the myokines IL-6 and irisin are important signals from the muscle that coordinate the beneficial adaptations to regular exercise such as weight loss and resistance to metabolic disease.

 

The Gene Expression Pathway that Mediates Muscle Adaptations to Exercise

One way to produce an adaptation within a tissue in response to a stimulus is to either activate or inhibit the gene expression pathway (Figure 14-9). This usually means that the stimulus has to initially modify gene transcription. The mRNA that results is translated into protein which contributes to changing the “phenotype” or appearance of the muscle cell. For example, if the transcription of MHC Type I is increased, then the muscle will become slower in its contractile properties. If the transcription of a mitochondrial lipid oxidation enzyme is increased, increased protein levels of the enzyme will cause the muscle to begin to favour lipid over COH metabolism.

What are typical stimuli that result in adaptations within skeletal muscle? Exercise is one, of course, but muscle also responds to conditions such as hypoxia, hibernation, heat and cold exposure, altitude, muscle disuse, and nutrient deprivation, to name a few. Indeed, muscle is considered to be highly “plastic”, meaning that it adapts readily to a number of external stimuli. In this chapter, we will consider only the adaptive response to exercise. While adaptations to any stimulus take time, you may be surprised to learn that even the first exercise bout can initiate the adaptation process.

There are several levels of “control” over this pathway, generally classified into

  • transcriptional; and
  • post-transcriptional regulation.

Transcriptional control is exerted within the nucleus, while post-transcriptional regulation can take place in the nucleus, cytoplasm or within organelles of the cell.

The Signals Involved in Triggering Muscle Adaptations to Exercise

The adaptation process begins with the response of muscle cells to the first seconds of an acute bout of exercise. Signals are transmitted from one protein to another via physical interactions which lead to changes in protein shape (i.e. conformational changes). Often these interactions arise because of the actions of an enzyme like a kinase that phosphorylates a protein or a phosphatase which removes a phosphate group (Figure 14-10). This is a very common method of transmitting a signal within a cell in response to a stimulus, like exercise, or hormone treatment, for example.

There are six possible signals to be considered that occur in muscle at the start of each exercise bout (Figure 14-11):

Voltage-sensitive proteins

When a motor unit is recruited and muscle cell membranes are depolarized, a voltage signal travels across the plasmalemma. Many proteins within the muscle membrane system are voltage-sensitive (e.g. Na+ channels, receptors) and their conformation is altered in response to differences in voltage. When that happens, they can often interact with other proteins to begin a signaling cascade.

The sequence of events that alter muscle behaviour or observable properties are listed in a flow diagram. At the top, a "Signal" will either activate or inactive DNA, resulting in a change in transcription, leading to a change in messenger RNA. Then translation is needed to generate protein. The new protein will give the muscle a different behaviour or phenotype.
Figure 14-9 Exercise produces intracellular “signals” which alter the gene expression pathway within myonuclei. These signals promote selective gene transcription, and the mRNA product is then translated into protein. The change in specific protein concentration within the muscle cell alters the appearance and function of the muscle fibre (i.e. changes the fiber phenotype).
A common way that the body recognizes or remembers a signal is phosphorylation of an enzyme. Here, phosphorylation is shown to result from an intracellular signal (first event in the image) activating a kinase, which takes the terminal phosphate from ATP and adds it to an inactive protein (purple circle) which then becomes active (purple box with red "P". At the bottom of the diagram, A phosphatase is shown to reverse the reaction, removing the phosphate from the active protein, making it inactive.
Figure 14-10 The first step in signal transduction in response to exercise is often via the activation of specific proteins such as transcription factors by phosphorylation. Repression of transcription factor activity can be achieved via de-phosphorylation. The enzymes responsible for this are kinases and phosphatases, respectively
This is an elegant and complex cartoon that illustrates many ways that gene expression can be altered. In the lower left corner, is part of a nucleus, with a strand of DNA, representing the ability to synthesize a wide array of proteins in the myocyte. There are six arrows pointing at a purple box with the text "activation of gene expression" just inside the nuclear envelope. 1. a mitochondrion is drawn in the lower left, depicting ATP production by oxidative metabolism. Mitochondrial activity will result in the release of reactive oxygen species and these can activate gene expression. 2. ion channels are represented at the membrane of the myocyte. Voltage sensitive proteins can respond to changes in ion distribution that will alter the membrane potential. 3. a neuromuscular junction comes in from the left, and is positioned just above the end plate. Activation of the muscle leads to Ca2+ release and Ca2+ not only activates the muscle but can directly activate gene expression, or indirectly do so via activation of an enzyme, calmodulin kinase. 4. At the top left is a box describing "change in muscle shape", with an arrow pointing to an integral membrane protein that is sensitive to distortion and signals activation of gene expression. 5. a couple of membrane embedded receptors are shown, binding to small molecules that have been released from the myocyte, allowing a self-signalling process. The release can be by exocytosis, which is also illustrated with a couple of vesicles inside the myocyte and one in the process of releasing the autocrine molecule. The binding of the signaling molecule to the receptor can activate gene expression. 6. Metabolic activity, particularly cross-bridge activity, illustrated by a sarcomere, labeled contracion aparatus can also activate gene expression. Here, contractile activity leads to hydrolysis of ATP and a small increase in ADP concentration. This activates AMPK which controls activation of gene expression.
Figure 14-11 Summary of the cellular events occurring during muscle contraction which could activate signals leading to changes in the gene expression pathway. See the text for details.
Mechanosensitive proteins

When a muscle cell contracts, a number of proteins are activated in response to changes in the shape of the muscle cell. In other words, they respond to mechanical deformation and trigger further protein activation “downstream”.

Cytosolic calcium

The propagation of the action potential causes the release of calcium from the SR. While calcium is important for mediating the contraction process, it is also a powerful activator of a number of kinase and phosphatase enzymes that can modify proteins by phosphorylation or dephosphorylation.

ATP turnover and the activation of AMP kinase

When calcium binds to troponin and permits and myosin to interact, is broken down to ADP and some ADP is further converted to AMP. AMP is an allosteric activator of the enzyme AMP kinase (hence its name!) which phosphorylates a number of transcription factors or other proteins which regulate transcription processes within the nucleus by binding to DNA in a sequence-specific manner.

Reactive O2 Species (ROS) production

ADP  is the most important activator of phosphorylation (i.e. O2 consumption and production in the mitochondria). When O2 consumption is increased, a small percentage of the O2 forms reactive oxygen species (ROS) instead of being converted to water. These molecules can be damaging when present in excess or beneficial when produced in moderate amounts since they can serve to oxidize and activate proteins.

Secreted proteins acting in autocrine fashion

When a muscle cell contracts, a number of myokines are secreted from the muscle cells into the extracellular space. These molecules may act on specific receptors on the surface of the muscle cell which secreted them, termed autocrine signaling. The binding of the myokine to its receptor triggers a series of signaling events, usually phosphorylation reactions, which activate proteins to modify gene expression. An excellent example of this is the secretion of insulin-like growth factor-1 (IGF-1) from muscle cells, which then triggers protein synthesis in muscle cells. Some newly discovered myokines that are released upon exercise include interleukin-6 (IL-6) and irisin. These circulate in the blood and can have metabolic effects in other tissues. Thus, muscle not only produces molecules that have autocrine effects, but can act as an endocrine organ as well! (See the Highlight Box for additional details).

Gene Expression and Splicing

The activation of proteins by any one of these six mechanisms can lead to a change in their shape and location within the cell. In the case of proteins such as transcription factors, random movements within the cell can allow them to be translocated from the cytosol to the nucleus. There, these transcription factors bind DNA within the control region of genes in a sequence-specific manner. The particular sequence to which a transcription factor binds is called the response element or transcription factor-binding site (Figure 14.12). Recall that the structure of genes includes DNA which is “coding” (i.e. exons) and that which is non-coding (introns). Upstream of these gene segments is the control region called the promoter of the genes. It is usually within this upstream region that transcription factors bind to either facilitate or inhibit, the activity of RNA polymerase, the enzyme that actually transcribes the DNA into mRNA. Additional proteins such as co-activators and co-repressors work in tandem with the transcription factor to activate or repress the expression of the gene of interest. Co-activators and co-repressors do not directly bind to the DNA, but instead interact with the transcription factor itself to affect its action on gene expression. Thus, the ability of the signals described above to activate or inhibit transcription factors, and the consequent effects on genes within the nucleus, are the key elements of transcriptional control of the gene expression pathway.

A large nucleus is shown in the diagram. At the. top is a segment of DNA, with a small part of it blown up across the bottom. The segment has: response element and muscle protein coding genes. The response element has transcription factor bound to it, and it is indicated that exercise signals with positive and negative impact on these transcription factors by bound coactivator or corepressor molecules also shown bound to the transcription factor. The transcription factor is also bound to RNA polymerase, which initiates gene transcription. Nuclear-respiratory factor-1 is given as an example of an exercise related
Figure 14-12 Exercise signals gene transcription.  At the level of the DNA within each myonucleus, cellular signals provoked by exercise lead to the binding of transcription factors (TF) to the response element (DNA) upstream of the coding region of a gene. The action of the TF is often aided or inhibited by coactivator (CoA) or corepressor (CoR) proteins, respectively. TF activity facilitates RNA Polymerase (RNA Pol) to transcribe the gene into mRNA for subsequent translation. Nuclear-respiratory factor-1 (NRF-1) is a typical protein which helps promote the synthesis of in muscle in response to repeated exercise bouts.

Once a gene has been transcribed, the mRNA is referred to as the primary transcript. This transcript still contains intronic regions of the gene and is subject to the process of splicing within the nucleus to remove these non-coding regions. This activity is controlled by splicing proteins. Different cell types (e.g. liver or heart or muscle) can splice a primary transcript differently depending on the activity of these splicing proteins. This can lead to mature mRNA sequences which differ from one another in different tissues, thereby coding for tissue-specific variations in the protein derived from that gene. This is a mechanism of producing divergent isoforms of proteins as discussed earlier in the chapter. Thus, one gene does not always encode one protein; different isoforms can represent more than one gene product from the same gene. Some isoforms arise from different genes (e.g. myosin heavy chain genes) while others can be derived from the same gene. An example of this is the fast myosin light chain gene within muscle (Figure 14-13).

mRNA Stability and Translation

The mature mRNA, once formed in the nucleus, is exported via nuclear pores into the cytoplasm where it can follow a number of paths. It can be protected, stabilized and then translated into protein. Alternately, it can be subject to degradation by RNAses within the cytoplasm. Thus, the fate of an mRNA molecule is a result of competition between stabilizing and degradative factors in the cytoplasm of the cell (Figure 14-14). Assuming that the translation process wins out, the initiation of translation is the “rate-limiting” step in the formation of a protein. Translation initiation is mediated by factors which require activation by signals similar to those described above for transcriptional activation. It involves the assembly of the ribosomal subunits around the mRNA and the start of ribosome movement down the mRNA molecule. The ribosome “reads” the mRNA one triplet codon at a time and permits the insertion of charged tRNAs to enter with their corresponding amino acid at each mRNA codon. This is the elongation phase of translation. Ribosomal proteins then assemble the growing polypeptide chain, one amino acid at a time, as the ribosome moves along the mRNA. When the termination codon is reached, no new tRNA is attached, and the polypeptide chain is released as a newborn protein molecule.

Protein Trafficking and Assembly

New proteins can be inserted as they are made into the endoplasmic reticulum (ER) of cells for further modification. This is called co-translational trafficking. The ER contains enzymes which can post-translationally modify proteins to increase their activity. These modifications include the cleaving of inactive regions of the protein or adding COH residues to the protein (i.e. a glycosylation reaction), to name a few. Alternatively, the newly-made protein can be bound by chaperone proteins which serve to traffic the protein to its correct location. An example of this is the movement of a protein from the cytoplasm to the mitochondrion (see below). Chaperones are bound to this newly synthesized protein at specific locations within the polypeptide sequence and they serve to target the protein to its correct location based on the amino acid sequence found within it. Thus, a chaperone acts like a taxi cab, serving to direct a protein to its correct location within the cell while the “driver” of the cab is the targeting sequence found on the protein which tells the taxi, or chaperone, where to go.

Examples of Exercise-Induced Adaptations in Muscle

This diagram illustrates the location of genes encoding the light chains of myosin on Chromosome 16. A strand of DNA is shown in black with a light brown segment. This light brown segment is expanded in the bottom of the image, showing alternative splicing yields mRNA1 coding for essential myosin light chains and mRNA2 coding for regulatory light chains.
Figure 14-13 Arrangements of the genes encoding MHC isoforms (Top) and myosin light chains (MLC; Bottom). MHC isoforms are encoded by different genes on chromosome 17, whereas MLC isoforms are generated by transcription of the same gene, but the mRNA product are spliced in different ways to produce two different proteins.
An oblong light brown cell is shown, with a round nucleus shown in blue on the left. A short arrow shows a strand of mRNA derived from a segment of DNA. Another arrow indicates the emergence of this mRNA from the nucleus into the cytoplasm, where it can follow 1 of 2 outcomes. It meets up with an enzyme, RNAse that leads to degradation, or it meets up with ribosomes that undergo translation, resulting in the synthesis of protein.
Figure 14-14 mRNA molecules emerging from the nucleus following transcription can be translated into protein by the action of ribosomes, or they can be degraded by the interaction of RNAses within the cytoplasm. The length of time that an mRNA molecule remains intact in the cytoplasm prior to degradation can determine how much protein is translated from the specific mRNA molecule.

Myosin Heavy Chain Isoform (MHC) Expression and Muscle Fibre-Type Conversions

As noted earlier, myosin heavy chains exist in several predominant isoforms within muscle. These include the MHC Type I, IIa and IIx, and they are so abundant in muscle that their expression determines the muscle fibre types of the same name. MHCs are large proteins of approximately 200 kDa in size. They are encoded by separate genes within the genome which lie in tandem beside each other on chromosome 17 (Figure 14.13). The factors that determine the expression of these genes are of great interest to exercise physiologists and fitness buffs alike. If we knew how to activate the expression of these genes, we could potentially switch fibre types and improve performance, or muscle health, at will.

What is the evidence that exercise can produce an adaptation in MHC gene expression, and therefore in muscle fibre types? There is plenty of evidence, but the result may not be what you expect. Recall that there are three main muscle fibre types, Type I (slow-twitch), Type IIa (fast-twitch) and Type IIx (also fast-twitch). These are so named because of the abundance of MHC isoforms (of the same name) within these fibres as detected by histochemical staining. In order for a fibre type “switch” to occur as result of training, the expression of the predominant MHC isoform within that fibre has to be reduced while that of another MHC isoform has to be increased.

Evidence from histochemical studies reveals that both endurance and resistance exercise training result in the same fibre type shift. The abundance of the Type IIx fibres is reduced and that of Type IIa fibres increases (Figure 14-15). There is very little evidence for conversion of fast- to slow-twitch fibres with the regular training regimens imposed on human muscle, though a few exceptions to this general rule have been documented in the literature. That being said, plenty of evidence for a slow-to-fast fibre type conversion exists in other mammals subjected to artificially-imposed chronic electrical stimulation for up to 24 hours per day. In those cases, the workload that is imposed on the muscle is much higher than we, as humans, impose on our muscles. In those cases, the result is a successful conversion to slow muscle fibres due to a change in the expression of the MHC genes. This suggests that it may ultimately be possible in humans as well. Note that opposite conversion, that of Type IIa to Type IIx fibres, can occur in humans, and this is generally observed under conditions of chronic muscle disuse such as that brought about by immobilization, denervation, or prolonged inactivity.

Why is it then that superb, elite endurance athletes (e.g. high level marathoners) have a high proportion of slow-twitch fibres, and that elite performers (e.g. Olympic-caliber sprinters) have a large fraction of fast-twitch fibres in their locomotory muscles? These endowments were not brought about by training but are a result of “genetics”. While these athletes have the same complement of genes as normal healthy non-athletes, the expression of those genes has differed for those athletes ever since they were born. These athletes have a different pattern of gene expression than most other individuals either because of

  • subtle differences in the sequence of DNA that controls gene expression (i.e. in the promoter region); or
  • differences in the sensitivity of the gene expression pathway, at any of the levels that we have discussed above, to intracellular signals.

This was likely inherited to a large degree and modified slightly by training and environmental influences. Hence, we refer to this difference as “genetic”.

Mitochondrial Biogenesis and Endurance Performance

Early findings

In the late 1960s, John Holloszy and his colleagues at Washington University in St. Louis were the first to show that endurance training produced an increase in mitochondrial content within muscles that had been recruited during the training period. This pioneering study led to an outpouring of research in this field, and it is now a well-known fact that the mitochondrial adaptations to training have a large number of performance benefits. These include a better use of lipid as a fuel during long-term exercise and the sparing of COH stores as well as a lesser production of lactic acid, leading to an improved endurance performance.

There are two parts to this figure, both illustrating the expected change in firbre-type with training or inactivity. At the top of the figure is a sequence from myosin heavy chain IIx to myosin heavy chain IIa to myosin heavy chain I. Arrows between these terms go both ways, indicating a reversible change. At the top of the figure it states "Mechanical overload, Exercise" with a green block arrow beneath, indicating change to the right. However, the change from myosin heavy chain IIa to I is lighter and the arrow is outlined with a dashed line to indicate this transition is rare or very slow. Beneath these terms is a red arrow indicating a leftward shift, again with the part involving type I myosin heavy chain marked with lighter colour and dashed line around the arrow. This direction of change in fibre-type is the result of immobilization, exposure to microgravity and detraining. The lower part of the figure presents a histogram with changes in myosin heavy chain resulting from (left to right) endurance training, resistance training and exceptions. In all cases, type IIa increases by 5-6% and there is a corresponding decrease in type IIx (or b, for study of small rodents). The exceptions indicate increases in type I and are marked by dashed lines around the column.
Figure 14-15 Typical (and exceptional) changes in muscle fibre type determined by histochemistry as a result of chronic muscle use (eg. exercise). The vast majority of studies illustrate that both resistance and endurance training program produce similar changes in muscle fibre type, with an increase in Type IIa, and a decrease in the proportion of Type IIb fibers. Type I fibre proportion is usually unchanged, but a few studies (Exceptions) report modest increases in Type I fibers in response to endurance training. In response to chronic muscle disuse such as microgravity or immobilization, most research indicates that the fibre conversion is the opposite to that of training, with increases in Type IIx, and decreases in Type IIa fibres. The dashed lines in the arrowheads indicate the rarity of the fibre type conversion involving Type I fibres.
Molecular basis for these changes

Figure 14-16 illustrates the complexity of the process of mitochondrial biogenesis produced by exercise. This process requires the coordinated synthesis of over 1500 proteins to expand the mitochondrial network within muscle cells. Mitochondria are unique in that they possess their own DNA, termed mitochondrial DNA (mtDNA). This genome is very small, encoding only 13 proteins. Thus, the remainder (about 1487 proteins!) must be coded for by nuclear DNA and synthesized in the cytoplasm via translation and then imported via specific channels into the mitochondrion. These imported proteins include Krebs’ cycle enzymes, proteins within the electron transport chain (ETC), and many others that the organelle requires to make ATP. Imported proteins, along with those derived from mtDNA, are then assembled to form multi-subunit complexes which are required for the to function. Exercise is known to increase the transcription of nuclear genes encoding mitochondrial proteins (NuGEMPS), and to increase the transport of newly synthesized proteins into the mitochondrion. Exercise also increases the number of copies of mtDNA within the organelle, thereby increasing the capacity for mitochondrial biogenesis.

How does exercise do this? The signals evoked by exercise provoke an increase in the expression and activity of PGC-1α (Figure 14-16). PGC-1α is considered to be a “master” regulator of mitochondrial synthesis. It facilitates the transcriptional activity of a large number of transcription factors in the nucleus including Nuclear respiratory factor-1 (NRF-1). NRF-1 transcribes a whole host of NuGEMPs, and the mRNAs encoded then escape the nucleus and are translated into protein in the cytoplasm. Some examples of these proteins include cytochrome c and subunits of the electron transport chain.

The importance of PGC-1α in mediating these adaptations in muscle is evident from animal studies in which mice are genetically engineered to have too much PGC-1α (i.e. Transgenic overexpressing animals), or too little (i.e. Knockout mice). These types of genetic manipulations in animals have proven to be extremely useful in medical science research. In the transgenic overexpressing animals, the muscle mitochondrial content is much higher than normal and so is the endurance performance. In contrast, in the knockout animals’ mitochondrial content is reduced and performance is impaired. Thus, it is clear that an understanding of how PGC-1α works in the cell can have a big impact on our knowledge of muscle adaptations.

Muscle Hypertrophy

Muscles enlarge in our bodies by undergoing hypertrophy as a result of an increase in the size of each individual muscle cell. Muscles typically do not divide or multiply, a process termed hyperplasia -they simply enlarge. The growth of individual muscle cells commonly occurs with resistance training that is performed with an appropriate intensity and duration as discussed earlier. The hypertrophy occurs as a result of two major types of events:

  • anabolic processes; and
  • satellite cell activation, fusion and donation of new myonuclei.
This cartoon illustrates the sequence of events for some forms of mitochondrial protein synthesis, triggered by chronic exercise. In this case, a beige nucleus is shown in the top left and a yellow mitochondrion is shown in the bottom right. Within the nucleus, transcription factor NRF-1 triggers the production of mRNA, which escapes from the nucleus and interacts with ribosomes in the cytoplasm to synthesize precursor proteins. The precursor proteins are taken up into the mitochondria, where they can be combined with other peptides, to form mature proteins. The mature proteins can be assembled with proteins synthesized within the mitochondria, from mitochondrial DNA and inserted into the mitochondrial membrane
Figure 14-16 Mitochondrial biogenesis in muscle occurs as a result of chronic exercise. First, gene are transcribed in the nucleus via the action of transcription factors like NRF-1. The mRNA is translated in the cytoplasm into a precursor protein with a positively charged targeting sequence which helps direct it to the mitochondrion import channel. The protein is imported into the organelle, the targeting sequence is cleaved, and the mature protein is refolded to act as an enzyme or assembled as a component of the electron transport chain.

The result is an increase in the number of contractile myofibrils as a consequence of increased expression of myosin, and regulatory protein genes that comprise the majority of the proteins within muscle cells. This muscular hypertrophy causes myofibrils to thicken and then divide, making more myofibrils. The increase in mass and cross-sectional area of a muscle is directly proportional to its ability to produce force.

Anabolic Pathways

These pathways are activated by resistance exercise and lead to an increase in protein synthesis. They involve a number of important signaling proteins. One of these proteins, called mTOR (mammalian target of rapamycin), plays a crucial role in the formation of new proteins. The mTOR protein is a kinase which phosphorylates and activates proteins such as S6Kinase and 4E-BP1 to activate translation (Figure 14-17). Calcium is also an essential player for skeletal muscle hypertrophy. With the release of calcium from the SR with every muscle contraction, calcium can activate a protein phosphatase called calcineurin. In contrast to mTOR, calcineurin removes phosphate groups from other proteins, particularly the nuclear factor of activated T-cells (NFAT), a name derived from the cell type in which it was originally discovered. When the phosphate group is removed from NFAT, it becomes free to enter the nucleus and increase the transcriptional activation of genes important for muscle growth like α- and IGF-1.

Satellite Cell Activation

Satellite cells (SC) are normally small dormant cells which reside adjacent to and within the basal lamina but outside of the plasmalemma of the much larger muscle cells (Figure 14.18). When skeletal muscle is exposed to an overload stimulus, factors such as Insulin-like Growth Factor-1 GF-1.are released which act on the neighboring satellite cells, causing them to become activated as well. Once activated, SC can divide to form “daughter” SC or fuse to the mature muscle cell and donate their nuclei and cytoplasm to that much larger muscle cell. Thus, the number of nuclei within a muscle cell can be altered by resistance training. This fusion and donation of nuclei by SC is an important adaptation which leads to muscle cell hypertrophy in response to this form of training. In fact, experiments have shown that if neighboring SC are destroyed prior to the start of training, very little hypertrophy will occur! It stands to reason that if a muscle cell gains a myonucleus and that nucleus continues to be active in gene expression, then the gene products (i.e. proteins) will contribute meaningfully to the size of the cell.

In contrast to muscle growth and hypertrophy of muscle fibres, disuse conditions such as casting, space flight, loss of innervation, or prolonged bed rest produce atrophy of fibres and of the whole muscle. This is likely due to a combination of a decrease in myofibrillar protein synthesis and an increase in protein degradation. In these situations, the number of myonuclei remains constant while the muscle fibre shrinks (Figure 14-18). Maintaining the same number of nuclei in light of a decrease in muscle mass allows for more rapid recovery. This may help explain the phenomenon of muscle “memory”, in which re-adaptations following injury or disuse are more rapid than they were originally when training first began. It should be noted that this is a highly studied area of research, and some controversy still exists in the field.

Research Box: Myonuclear Domain Theory

The myonuclear domain (MND) theory, originally conceptualized by Cheek and colleagues in 1971 (1985), has been widely accepted in research until very recently when it gathered a storm of controversy. Simply put, the theory postulates that each nucleus in a myofibre is responsible for a theoretical amount of cytoplasm and that this nuclear-to-cytoplasmic ratio adapts to changes in muscle size. During muscle hypertrophy, the muscle grows in size either by increasing the cytoplasmic domain size per nucleus or by augmenting the number of myonuclei thereby keeping the MND size constant. Conversely, the theory posits that under conditions ofatrophy as the myonuclei are targeted for destruction, there is an accompanying yet relatively greater loss of cytoplasmic volume and the MND size is reduced.

Much of the work in the field has been conducted by Edgerton and colleagues (1991) who have rigorously studied myonuclear number and domain size during muscle hypertrophy and atrophy  Earliest reports suggested an elevation in the number of myonuclei commensurate with an increase in fibre size during muscle hypertrophy wherein the nuclear-to-cytoplasmic ratio remains unchanged. However, it was later illustrated that increases in muscle size can occur independently of changes in nuclei count. In such cases, the domain size has been proposed to increase to a 2000μm threshold ceiling subsequent to which an increase in nuclei number is necessary for further muscle hypertrophy (Allen, 1999). Furthermore, MNDs increase in size during maturational muscle growth and are fibre-type dependent, with the smallest domains found in type I and the largest in type IIB/X fibres. Thus, the literature supports the existence of a constant or increasing size of nuclear domains with muscle hypertrophy.

A multitude of studies have suggested that myonuclei are targeted for degradation and the subsequent loss of MNDs explains the observed muscle atrophy during conditions such as disuse, denervation, neuromuscular disorders and mechanical unloading (Bruusgaard et al., 2012). This notion was recently challenged vigorously by Gundersen and colleagues (Bruusgaard and Gundersen 2008, Gunderson, 2011) who illustrated that these studies had failed to distinguish between the muscle nuclei and nuclei from surrounding cells undergoing apoptosis such as stromal and satellite cells. Using time lapse imaging of a single fibre in vivo, they illustrated no loss of myonuclei despite a 50% reduction in fibre size. Even during aging, a clear pattern has emerged that myonuclei are not lost despite substantial atrophy. In fact, some studies even report an increase in myonuclei per muscle fibre (Petrella, et al., 2006). Clearly, the data are in direct opposition to the MND theory as they reflect that muscle atrophy occurs independently of the loss of myonuclei and is therefore irrespective of the principle of a nuclear-based control over cytoplasmic domains.

Despite the evidence presented by Gundersen et al. (Bruusgaard and Gundersen 2008, Gunderson, 2011) in recent years that challenges the MND theory, it would be premature and unwise to discount the plethora of studies published over the last 30 years that show otherwise. Differences in methodology and study design can account for some of the variability observed. Clearly, this area of research certainly warrants the attention it is currently receiving and may result in a universally agreed upon adjustment to the MND theory and its role in muscle size.

Summary

Muscles are composed of different fibre types which adapt differently depending on the type of exercise imposed. Resistance training typically leads to muscle fibre hypertrophy, thereby increasing strength, while endurance training leads to metabolic adaptations in mitochondrial function, producing greater endurance. The bases for these adaptations reside in exercise-induced modifications of the gene expression pathway, consisting of signaling, gene transcription, mRNA processing, translation, and post-translational modifications and trafficking. A single acute bout of exercise can initiate the adaptation process by activating this pathway. Repeated bouts of exercise (i.e. training) can lead to the accumulation of gene products (i.e. proteins) that result in an alteration in muscle phenotype, and improved function. Understanding how the gene expression pathway works and how exercise can modify it is the basis for our comprehension of exercise adaptations.

This figure summarizes the sequence of events resulting from resistance exercise. A segment of a muscle fibre is shown with initial event "resistance exercise across the top. Arrows indicate 2 consequences of this exercise: release of Ca2+ from the sarcoplasmic reticulum and increased mTOR (mammalian target of rapamycin). Sarcoplasmic reticulum is drawn on the right, with Ca2+ indicated as being released into the cytoplasm. An arrow from Ca2+ to CaN (calcineurin) in an orange box indicates activation sequence. The next arrow points at the change from phosphorylated NFAT to unphosphorylated NFAT. Calcineurin is a phosphatase. Unphosphorylated NFAT is indicated as migrating into the nucleus, a blue oval in the top area of the fibre segment. Following the arrows, this results in release of mRNAs from the nucleus into the cytoplasm. In the cytoplasm, ribosomes are drawn in blue. The mRNAs stimulate translation and protein synthesis. Meanwhile, the other arrow from "resistance exercise" at the top of the figure goes to mTOR inside the fibre, indicating release and an increase in concentration. mTOR has arrows going to cell signalling molecules: S6K and 4E-BP1, both of which have arrows going to the ribosomes, indicating regulation of translation leading to protein synthesis of actin and myosin needed for muscle cell hypertrophy.
Figure 14-17 Resistance exercise activates two pathways which signal muscle growth (or hypertrophy). First, the kinase mTOR is activated to phosphorylate two proteins required for translation: S6K and 4E-BP1. Once activated these proteins trigger the process of protein synthesis, including and myosin. In addition, contraction causes the release of calcium which binds to calcineurin (CaN) to dephosphorylate the transcription factor nuclear factor of activated T cells (NF-AT). This protein promotes the transcription of the and myosin genes to provide the mRNA for translation.
This figure shows a series of muscle fibre segments and the nuclie and satellite cells. The original myofibre is shown on the left; several nuclei are shown in the periphery with corresponding nuclear domain indicated by dashed lines around each nucleus. A couple of satellite cells are position beneath the basement membrane, but outside the cell plasmalemma. This fibre is labelled "Control". A large blue arrow points from this cell to a thinner cell above, labelled as "Atrophy". A similar number of nucleii are present, so the nuclear domain is smaller. Another blue arrow points from the control fibre to another fibre in the lower segment of the image. This fibre is undergoing hypertrophy. An early event associated with hypertrophy is proliferation of satellite cells, indicated by more cells between the myocyte plasmalemma and the basement membrane. then an arrow points from this cell to an adjacent larger fibre with more nuclei and a nuclear domain similar to the control cell. The number of satellite cells has decreased, indicating migration of these cells into the myobibre.
Figure 14-18 Muscle fibers atrophy in response to chronic disuse, and hypertrophy as a result of resistance training. See text for details. Adapted from Saleem A. et al., in Reed J. and D. Green, Eds. Apoptosis: Physiology and Pathology. New York: Cambridge, 2011, p. 318.


References

Adams GR, Hather BM, Baldwin KM, and Dudley GA. 1993. Skeletal Muscle Myosin Heavy Chain Composition and Resistance Training. Journal of Applied Physiology 74(2):911-915.

Allen DL, Roy RR, and Edgerton VR. 1999. Myonuclear domains in muscle adaptation and disease. Muscle & Nerve 22(10):1350-60. doi:10.1002/(SICI)1097-4598(199910)22:10<1350::AID-MUS3>3.0.CO;2-8.

Astorino TA, Allen RP, Roberson DW, Jurancich M, Lewis R, McCarthy K, and Trost E. 2011. Adaptations to high-intensity training are independent of gender. European Journal of Applied Physiology 111(7):1279-86. doi:10.1007/s00421-010-1741-y.

Barton ER. 2006. The ABCs of IGF-I isoforms: impact on muscle hypertrophy and implications for repair. Applied Physiology, Nutrition, and Metabolism 31(6):791-7. doi:10.1139/h06-054.

Baumann H, Jäggi M, Soland F, Howald H, and Schaub MC. 1987. Exercise training induces transitions of myosin isoform subunits within histochemically typed human muscle fibres. Pflügers Archiv : European Journal of Physiology 409(4-5):349-60. doi:10.1007/BF00583788.

Berchtold MW, Brinkmeier H, and Müntener M. 2000. Calcium ion in skeletal muscle: its crucial role for muscle function, plasticity, and disease. Physiological Reviews 80(3):1215-65. doi: 10.1152/physrev.2000.80.3.1215

Biressi S, and Rando TA. 2010. Heterogeneity in the muscle satellite cell population. Seminars in Cell & Developmental Biology 21(8):845-54. doi:10.1016/j.semcdb.2010.09.003.

Bortoluzzi S, Scannapieco P, Cestaro A, Danieli GA, and Schiaffino S. 2006. Computational reconstruction of the human skeletal muscle secretome. Proteins 62(3):776-92. doi:10.1002/prot.20803.

Boström P, Wu J, Jedrychowski MP, Korde A, Ye L, Lo JC, Rasbach KA, Boström EA, Choi JH, Long JZ, Kajimura S, Zingaretti MC, Vind BF, Tu H, Cinti S, Højlund K, Gygi SP, and Spiegelman BM. 2012. A PGC1-α-dependent myokine that drives brown-fat-like of white and thermogenesis. Nature 481(7382):463-8. doi:10.1038/nature10777

Bruusgaard JC, Egner IM, Larsen TK, Dupre-Aucouturier S, Desplanches D, and Gundersen K. 2012. No change in myonuclear number during muscle unloading and reloading. Journal of Applied Physiology 113(2):290-6. doi:10.1152/japplphysiol.00436.2012.

Bruusgaard JC, and Gundersen K. 2008. In vivo time-lapse microscopy reveals no loss of murine myonuclei during weeks of muscle atrophy. The Journal of Clinical Investigation 118(4):1450-7. doi:10.1172/JCI34022.

Bruusgaard JC, Johansen IB, Egner IM, Rana ZA, and Gundersen K. 2010. Myonuclei acquired by overload exercise precede hypertrophy and are not lost on detraining. Proceedings of the National Academy of Science of the United States of America. 107(34):15111-15116. doi:10.1073/pnas.0913935107.

Calvo JA, Daniels TG, Wang X, Paul A, Lin J, Spiegelman BM, Stevenson SC, and Rangwala SM. 2008. Muscle-specific expression of PPARgamma coactivator-1alpha improves exercise performance and increases peak oxygen uptake. Journal of Applied Physiology 104(5):1304-12. doi:10.1152/japplphysiol.01231.2007.

Cheek D. 1985. The control of cell mass and replication. The DNA unit-a personal 20-year study. Early Human Development 12(3):211-239. doi:10.1016/0378-3782(85)90144-6.

Crawford GL, and Horowits R. 2011. Scaffolds and chaperones in assembly: Putting the striations in striated muscle. Biophysical Reviews 3(1):25-32. doi:10.1007/s12551-011-0043-x.

Edgerton VR, and Roy RR. 1991. Regulation of skeletal muscle fiber size, shape and function. Journal of Biomechanics 24 Suppl 1:123-133. doi:10.1016/0021-9290(91)90383-X.

Florini JR, Magri KA, Ewton DZ, James PL, Grindstaff K, and Rotwein PS. 1991. “Spontaneous” differentiation of skeletal myoblasts is dependent upon autocrine secretion of insulin-like growth factor-II. The Journal of Biological Chemistry 266(24):15917-23.

Green HJ, Klug GA, Reichmann H, Seedorf U, Wiehrer W, and Pette D. 1984. Exercise-induced fibre type transitions with regard to myosin, parvalbumin, and sarcoplasmic reticulum in muscles of the rat. Pflügers Archiv : European Journal of Physiology 400(4):432-438. doi:10.1007/BF00587545.

Gundersen K. 2011. Excitation-transcription coupling in skeletal muscle: The molecular pathways of exercise. Biological Reviews of the Cambridge Philosophical Society 86(3):564-600. doi:10.1111/j.1469-185X.2010.00161.x.

Handschin C, Chin S, Li P, Liu F, Maratos-Flier E, LeBrasseur NK, Yan Z, and Spiegelman BM. 2007. Skeletal muscle fiber-type switching, exercise intolerance, and myopathy in PGC-1α muscle-specific knock-out animals. Journal of Biological Chemistry. 282:30014-30021. doi:10.1074/jbc.M704817200.

Hawke TJ, and Garry DJ. 2001. Myogenic satellite cells: physiology to molecular biology. Journal of Applied Physiology (Bethesda, Md. 1985) 91(2):534-51. doi:10.1152/jappl.2001.91.2.534.

Hood D. 2009. Mechanisms of exercise-induced mitochondrial biogensis in skeletal muscle. Applied Physiology, Nutrition, and Metabolism 34(3):465-72. doi:10.1139/H09-049.

Hood DA. 2001. Invited Review: contractile activity-induced mitochondrial biogenesis biogenesis in skeletal muscle. Journal of Applied Physiology (Bethesda, Md. 1985) 90(3):1137-57. doi: 10.1152/jappl.2001.90.3.1137.

Hornberger TA. 2011. Mechanotransduction and the regulation of mTORC1 signaling in skeletal muscle. International Journal of Biochemistry & Cell Biology 43(9):1267-76. doi:10.1016/j.biocel.2011.05.007.

Liu D, Sartor MA, Nader GA, Gutmann L, Treutelaar MK, Pistilli EE, IglayReger HB, Burant CF, Hoffman EP, and Gordon PM. 2010. Skeletal muscle gene expression in response to resistance exercise: Sex specific regulation. BMC Genomics 11:659. doi:10.1186/1471-2164-11-659.

Ljubicic V, Joseph A-M, Saleem A, Uguccioni G, Collu-Marchese M, Lai RYJ, Nguyen LM-D, and Hood DA. 2010. Transcriptional and post-transcriptional regulation of mitochondrial biogenesis in skeletal muscle: Effects of exercise and aging. Biochimica et Biophysica Acta – General Subjects 1800(3):223-34. doi:10.1016/j.bbagen.2009.07.031.

Miller AE, MacDougall JD, Tarnopolsky MA, and Sale DG. 1993. Gender differences in strength and muscle fiber characteristics. European Journal of Applied Physiology and Occupational Physiology 66(3):254-62. doi:10.1007/BF00235103.

Ordway GA, and Garry DJ. 2004. Myoglobin: an essential hemoprotein in striated muscle. Journal of Experimental Biology 207:3441-3446. doi:10.1242/jeb.01172.

Pattengale PK, and Holloszy JO. 1967. Augmentation by a program of skeletal muscle myoglobin of treadmill running. American Journal of Physiology 213(3): 783–785.

Pedersen BK, and Febbraio MA. 2008. Muscle as an Endocrine Organ: Focus on Muscle-Derived Interleukin-6. Physiological Reviews 88(4):1379-406. doi:10.1152/physrev.90100.2007.

Petrella JK, Kim JS, Cross JM, Kosek DJ, and Bamman MM. 2006. Efficacy of myonuclear addition may explain differential myofiber growth among resistance-trained young and older men and women. American Journal of Physiology. Endocrinology and Metabolism 291(5):E937-46 doi:10.1152/ajpendo.00190.2006.

Roepstorff C, Thiele M, Hillig T, Pilegaard H, Richter EA, Wojtaszewski JFP, and Kiens B. 2006. Higher skeletal muscle alpha2AMPK activation and lower energy charge and oxidation in men than in women during submaximal exercise. Journal of Physiology 574(Pt 1):125-38. doi:10.1113/jphysiol.2006.108720.

Rusko HK. 1992. Development of aerobic in relation to age and training in cross-country skiers. Medicine and Science in Sports and Exercise 24(9):1040-7.

Sandri M. 2008. Signaling in Muscle Atrophy and Hypertrophy. Physiology 23:160-170. doi:10.1152/physiol.00041.2007.

Scarpulla RC. 2011. Metabolic control of mitochondrial biogenesis biogenesis through the PGC-1 family regulatory network. Biochimica et Biophysica Acta 1813(7):1269-78 doi:10.1016/j.bbamcr.2010.09.019.

Short KR, Vittone JL, Bigelow ML, Proctor DN, Coenen-Schimke JM, Rys P, and Nair KS. 2005. Changes in myosin heavy chain mRNA and protein expression in human skeletal muscle with age and endurance exercise training. Journal of Applied Physiology 99:95-102. doi:10.1152/japplphysiol.00129.2005.

Steensberg A, Keller C, Starkie RL, Osada T, Febbraio MA, and Pedersen BK. 2002. IL-6 and TNF-alpha express1`ion in, and release from, contracting human skeletal muscle. American Journal of Physiology. Endocrinology and Metabolism 283(6):E1272-8. doi:10.1152/ajpendo.00255.2002.

Van der Meer SFT, Jaspers RT, and Degens H. 2011. Is the myonuclear domain size fixed? Journal of Musculoskeletal & Neuronal Interactions 11(4):286-97.

Williamson DL, Godard MP, Porter DA, Costill DL, and Trappe SW. 2000. Progressive resistance training training reduces myosin heavy chain coexpression in single muscle fibers from older men. Journal of Applied Physiology (Bethesda, Md. : 1985) 88(2):627-33. doi:10.1152/jappl.2000.88.2.627.

West DWD, Burd NA, Churchward-Venne TA, Camera DM, Mitchell CJ, Baker SK, Hawley JA, Coffey VG, and Phillips SM. 2012. Sex-based comparisons of myofibrillar protein synthesis after resistance exercise in the fed state. Journal of Applied Physiology 112(11):1805-13. doi:10.1152/japplphysiol.00170.2012.

Wüst RCI, Morse CI, De Haan A, Jones DA, and Degens H. 2008. Sex differences in contractile properties and resistance of human skeletal muscle. Experimental Physiology 93(7):843-50. doi:10.1113/expphysiol.2007.041764.

Examine your adaptations to endurance training

Resources to have available:

  • Exercise facility with a calibrated bicycle ergometer or treadmill;
  • Portable blood lactate analyzer;
  • Heart rate monitor;
  • A partner to help with heart rate (HR) or breathing rate (f).

Purpose: to examine your individualized adaptations to endurance training

Methods:

  • Design your own endurance training regimen in which you exercise using a modality of your choice (e.g. Jogging, swimming, cycling) 3-4 times/week, at 60-90% of your maximum HR for at least 6 weeks. This program needs to be in addition to your current level of activity. Prior to starting this additional training, test your current fitness level using an indirect, submaximal fitness test (e.g. Astrand-Rhyming test). At a standardized, absolute workload, measure your HR and breathing frequency (f; here is where a partner could help). If possible, donate a finger prick blood drop for lactate analysis.
  • Complete the fitness training program.
  • Repeat your measurements again after 6 weeks of training, using the same test workload as in the pre-test.

Expected Results:

  • A lower HR and f are expected at the same absolute workload after training.
  • Reduced blood lactate concentrations are also expected, reflecting less lactic acid production by the working muscle. This is probably a result of an improved mitochondrial content (and therefore aerobic system energy provision) as an adaptation to training. This adaptation means that there is less reliance on anaerobic glycolysis for synthesis post-training, and greater endurance.

Introducing the Authors

image

D.A. Hood, , , , , School of Kinesiology and Health Science, Muscle Health Research Centre, York University, Toronto ON. David A. Hood is a Full Professor and Tier I Canada Research Chair in Cell Physiology in the School of Kinesiology and Health Science at York University. His research investigates the role of exercise and disuse on mitochondrial turnover in skeletal muscle.

 

 

 

 

 

 

N. Cheema, Harvard Medical School and Massachusetts General Hospital.

Nashwa Cheema is a postdoc in Dr. Rox Anderson’s lab in Harvard Medical School at Massachusetts General Hospital (HMS/MGH) and has been working on a project to enhance muscle performance and mitigate injuries with the application of photobiomodulation, the process of near-infrared light treatment. She obtained her PhD in 2017 at University of Alberta, Canada, in Physiology, Cell and Developmental Biology where her thesis was focused on sarcopenia, loss of muscle mass and function, in aged rats. Prior to joining HMS/MGH in 2020, Nashwa was a postdoc in Dr. David Hood’s laboratory at York University, Canada, a pioneer in the field of exercise physiology. Her project in his lab was focused on treatment strategies for mitochondrial health in isolated fibroblasts from patients with mitochondrial mutations.

 

Ayesha Saleem, Faculty of Kinesiology and Recreation Management, University of Manitoba, Children’s Hospital Research Institute of Manitoba (CHRIM), Manitoba

Ayesha Saleem, received her PhD degree from the School of Kinesiology and Health Science at York University in 2013.
She is now an Associate Professor at the Faculty of Kinesiology and Recreation Management, University of Manitoba, Winnipeg, Manitoba and a Research Scientist at CHRIM in Winnipeg, Manitoba where her laboratory is situated. Her research team studies extracellular vesicles in various models of health and disease, with a particular emphasis on the regulatory effect of extracellular vesicles on mitochondrial biogenesis.

 

 

 

 

 

Heather Carter –

image

Heather Carter received her PhD degree (2017) from the School of Kinesiology and Health Science at York University

She is now enrolled as a student at OISE, University of Toronto

 

 

 

 

 

 

Anna Vainshtein –Director, Craft Science Inc.

Dr. Anna Vainshtein is currently the director of Craft Science Inc., a science communications consulting firm based out of Toronto, Canada. She earned her B.Sc. (Hons), MSc. and Ph.D. at the School of Kinesiology and Health Science at York University, Toronto, Canada. During her academic journey, Dr. Vainshtein discovered and cultivated a profound passion for mitochondria, autophagy, muscle biology, and cellular bioenergetics. Seeking to deepen her understanding of mitophagy’s intricate roles in skeletal muscle, she pursued a fellowship at the Venetian Institute for Molecular Medicine in Padova, Italy. Subsequently, she completed her postdoctoral training at the Department of Molecular and Human Genetics at Baylor College of Medicine in Houston, Texas, USA, where she expanded her expertise to encompass the genetic regulation of the autophagy-lysosome system. Dr. Vainshtein’s journey reflects a relentless pursuit of knowledge and a commitment to advancing the frontiers of biomedical science on an international stage.

 

 

Olga Ostojic

Olga Green (née Ostojic) completed both a Bachelor’s (2010) and Master’s (2012) degree from the school of Kinesiology and Health Sciences at York University. She then went on to complete her Doctor of Dental Surgery degree at the University of Toronto. She partook in several dental outreach initiatives, including a rotation within a hospital-based clinic in Moose Factory, Ontario, where she worked with underserved First Nations populations. She now works as a general dentist in Toronto.

 

 

 

 

 

 

 

S. Iqbal

Sobia Iqbal – received her PhD degree (2015) from the School of Kinesiology and Health Science at York University

She is now a part-time faculty at Wilfrid Laurier University – Waterloo in the department of Biology

License

Icon for the Creative Commons Attribution-NonCommercial 4.0 International License

Open Textbook of Exercise Physiology Copyright © by David Hood; Nashwa Cheema; Ayesha Saleem; Heather Carter; Anna Vainshtein; Olga Ostojic; and Sobia Iqbal is licensed under a Creative Commons Attribution-NonCommercial 4.0 International License, except where otherwise noted.

Share This Book