You are on page 1of 23

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/11686440

Neural Adaptations to Resistance Training: Implications for Movement


Control

Article  in  Sports Medicine · February 2001


Source: PubMed

CITATIONS READS

157 5,799

3 authors, including:

Stephan Riek
University of the Sunshine Coast
167 PUBLICATIONS   4,410 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

excitability on visuo-motor adaptation View project

All content following this page was uploaded by Stephan Riek on 12 October 2014.

The user has requested enhancement of the downloaded file.


Acta Physiol 2011, 202, 119–140

REVIEW
Neural adaptations to strength training: Moving beyond
transcranial magnetic stimulation and reflex studies

T. J. Carroll,1 V. S. Selvanayagam,1,2 S. Riek1 and J. G. Semmler3

1 School of Human Movement Studies, The University of Queensland, Brisbane, Queensland, Australia
2 Sports Centre, University of Malaya, Kuala Lumpur, Malaysia
3 Discipline of Physiology, School of Medical Sciences, The University of Adelaide, Adelaide, South Australia, Australia

Received 4 October 2010, Abstract


revision requested 16 December It has long been believed that training for increased strength not only affects
2010,
muscle tissue, but also results in adaptive changes in the central nervous
revision received 12 January 2011,
accepted 2 March 2011 system. However, only in the last 10 years has the use of methods to study
Correspondence: Dr T. J. Carroll, the neurophysiological details of putative neural adaptations to training
Faculty of Health Sciences, School become widespread. There are now many published reports that have used
of Human Movement Studies, The single motor unit recordings, electrical stimulation of peripheral nerves, and
University of Queensland, Qld
non-invasive stimulation of the human brain [i.e. transcranial magnetic
4072, Australia.
E-mail: timothy.carroll@uq.edu.au
stimulation (TMS)] to study neural responses to strength training. In this
review, we aim to summarize what has been learned from single motor unit,
reflex and TMS studies, and identify the most promising avenues to advance
our conceptual understanding with these methods. We also consider the few
strength training studies that have employed alternative neurophysiological
techniques such as functional magnetic resonance imaging and electroen-
cephalography. The nature of the information that these techniques can
provide, as well as their major technical and conceptual pitfalls, are briefly
described. The overall conclusion of the review is that the current evidence
regarding neural adaptations to strength training is inconsistent and
incomplete. In order to move forward in our understanding, it will be nec-
essary to design studies that are based on a rigorous consideration of the
limitations of the available techniques, and that are specifically targeted to
address important conceptual questions.
Keywords H-reflex, muscle strength, single motor unit, transcranial
magnetic stimulation, V-wave.

There is a long history of research aimed at identifying hypertrophy. In the last 10 years or so, however, there
the nervous system responses to repetitive muscle has been a relative explosion in the use of neurophys-
actions against high loads (i.e. strength training). As iological methods to assess the effects of strength
an illustration, at least 25 systematic reviews have been training on various aspects of neural function. This
published in the last 30 years that have neural adapta- has occurred due the widespread availability of tech-
tions to strength training as a major focus. Until niques for non-invasive stimulation of the human brain
relatively recently, most of the evidence on the topic [i.e. transcranial magnetic stimulation (TMS)], as well
was indirect, such as the well-known mismatch between as electrical stimulators, amplifiers and associated
the time course of strength increases versus muscle devices for the capture and storage of electrical and

 2011 The Authors


Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x 119
Neural adaptations to strength training Æ T J Carroll et al. Acta Physiol 2011, 202, 119–140

twitch responses of muscle. The purpose of this review leaving the fine wires within the muscle, with the cut
is to critically evaluate evidence about the neural ends of the wires representing the exposed recording
responses to strength training obtained from single surface. More recently, concentric needle electrodes
motor unit, reflex, and TMS studies. The physiological have been used to record single motor unit activity,
and technical principles that underlie these techniques, which usually consist of a stiff needle (250 lm in
and their limitations and complications, are emphasized diameter) that remains within the muscle during the
in order to make the point that care is required when voluntary contraction (Stalberg et al. 1996). One
interpreting data from these electrophysiological meth- advantage of this technique is that the needle can be
ods in humans. We also consider the potential of some advanced (as opposed to only withdrawn) to optimize
alternate applications of these techniques to provide the motor unit recording, or manipulated to record
new information regarding the neural adaptations to from other locations within the muscle, which can allow
strength training. Subsequently, methods to assess recordings at high forces under some circumstances.
nervous system functions that have only very recently The obvious drawback is that the needle can generate
been applied to strength training will be considered. some discomfort within the muscle during a voluntary
These include techniques for studying brain function contraction, and it is not suitable for recording action
through functional magnetic resonance imaging (fMRI) potentials when there is a change in muscle length.
and electroencephalography (EEG). The overall goals Furthermore, recordings at high forces with this tech-
are to summarize the current evidence regarding neural nique are often limited by the inability to identify the
adaptations to strength training, and to identify the recruitment threshold or to obtain a sufficient number
most promising avenues to address the key conceptual of action potentials for each motor unit, because the
questions that remain. location of the electrode is often manipulated through-
out the contraction. Nonetheless, it seems that measur-
ing motor unit activity with a concentric needle
Single motor unit studies
electrode has become more commonplace over recent
Since its introduction in the late 1920s (Adrian & Bronk years, with a growing number of studies utilizing this
1929), intramuscular recording of single motor unit technique during isometric contractions.
activity has provided valuable insight into the neural In addition to intramuscular electrodes, motor unit
control of movement in humans. The key advantage of potentials can be recorded with subcutaneous electrodes
recording single motor unit activity is that the discharge or with surface electrode arrays. Subcutaneous elec-
properties of the motoneuron, whose cell body origi- trodes comprise a branched bipolar configuration that is
nates in the spinal cord, can be obtained from the positioned under the skin but over the muscle (Gydikov
discharge of the single motor unit. This is achieved et al. 1986, Enoka et al. 1988). Although more techni-
because every action potential generated in the moto- cally difficult to fabricate and use compared with
neuron will result in an action potential in all muscle conventional fine-wire electrodes, these electrodes are
fibres innervated by the parent motoneuron, because of capable of providing selective recordings of single
the high safety factor for action potential transmission motor unit potentials up to maximal contractions
at the neuromuscular junction (Bigland-Ritchie et al. (Oya et al. 2009), and provide the necessary stability
1979). Accordingly, single motor unit recording is one to isolate and identify the same motor unit during
of only a few neurophysiological techniques that can different fatiguing tasks (Miller et al. 1996, Mottram
provide unambiguous information about the behaviour et al. 2005). Electrode arrays consist of a grid of
of motoneurons during voluntary contractions in electrodes placed on the skin above the muscle, and
humans. were originally used to provide a non-invasive assess-
ment of the conduction velocity of motor unit potentials
along the muscle fibres (Schneider et al. 1989, see
Technical considerations
Merletti et al. 2003 for review). However, the develop-
The conventional approach to record single motor unit ment of more advanced processing methods has
activity is to insert an electrode into the muscle of expanded the utility of this technique through decom-
interest to record the extracellular action potentials position of the surface electromyogram (EMG) to
from a group of muscle fibres associated with one motor identify the discharge times of individual motor units
unit. An intramuscular electrode commonly consists of (De Luca et al. 2006, see Merletti et al. 2008 for
a set of flexible fine wires or a needle. A fine wire review), which can be accurate up to maximal contrac-
electrode is usually custom made and can consist of tions (Nawab et al. 2010). While these techniques are
several (2–4) insulated wires that are inserted into the limited to superficial motor units (Farina et al. 2010),
muscle with a hypodermic needle (Milner-Brown et al. they offer a promising alternative to more invasive
1973). With this technique the needle is withdrawn

 2011 The Authors


120 Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x
Acta Physiol 2011, 202, 119–140 T J Carroll et al. Æ Neural adaptations to strength training
procedures where interpretations are constrained must also be contributing to the early gains in strength
because of a low motor unit yield. with training.
Despite the emergence of some new hardware and Experimental evidence suggests that more subtle
software, longitudinal strength training studies pose an changes in single motor unit activity, such as an
additional challenge for reliable assessment of motor increased incidence of brief interspike intervals (known
unit activity in humans. The main problem is that it is as double discharges or doublets), may have the
not yet possible to identify the same motor unit before potential to influence the neural control of force with
and after the training program, in order to clearly strength training. When present at the beginning of a
demonstrate an effect of the intervention. Longitudinal voluntary contraction, double discharges have been
tracking of the same motor unit is possible under some shown to substantially increase low-frequency (unfused)
circumstances (Doherty & Brown 1994), but the motor unit force (Macefield et al. 1996), which may be
usefulness of the technique is likely to be limited to important in the maintenance of force during a
patient populations with substantial motoneuron loss submaximal fatiguing contraction (see Garland &
(Gooch & Harati 1997), and does not provide infor- Griffin 1999). After rapid force development training
mation on motor unit discharge behaviour during of the ankle dorsiflexors, Van Cutsem et al. (1998) has
voluntary contractions. Therefore, with current motor shown that an increase in the rate of force development
unit recording techniques, studies that involve strength (RFD) was associated with an increased incidence of
training require a large sample of motor units in each double discharges from 5% before training to 33% after
subject, with a sufficient number of subjects, to training. This finding suggests that it may be an
adequately characterize the change with training. appropriate neural strategy to utilize when performing
Because of these technical constraints, coupled with rapid muscle contractions. However, other evidence
the time consuming nature of the experiments, there are suggests that the implementation of double discharges
only a few studies that have adequately addressed the may not be a strategy used to improve the functional
effect of strength training on motor unit activity in outcome, but represent a byproduct of a change in the
humans. intrinsic properties of the motoneurons. For example,
the incidence of double discharges is not associated with
the speed of the voluntary contraction (Bawa &
Single motor unit activity
Calancie 1983, Christie & Kamen 2010), is reduced in
One possible neural mechanism to increase muscle older adults (Klass et al. 2008, Christie & Kamen
strength with training is an increase in motor unit 2010), and is greater in patients with neuromuscular
discharge rate during maximal contractions. Because disorders (see Thomas et al. 2002). The lack of any
these measures are technically challenging, only a consistent finding on the functional significance of
limited number of studies have examined maximal double discharges may be caused by sampling from
motor unit discharge rates after a strength training only a small number of motor units in each contraction,
intervention. Specific details of these studies are pro- and may depend on the muscle investigated, the
vided in Table 1. Using both isometric and dynamic contraction intensity, or the type of motor unit (low
training protocols that produced significant increases in vs. high threshold) examined (see Garland & Griffin
strength, these studies have demonstrated an increase in 1999, Christie & Kamen 2006).
maximal motor unit discharge rates in upper and lower Along with the gain in maximal strength, several
limb muscles after training, although this is not always lines of evidence suggest that training alters the activity
a consistent finding (e.g. Pucci et al. 2006). The greatest of single motor units and improves performance of
changes in maximal discharge rates with training (up to submaximal tasks. In particular, after several weeks of
49% increase) have been observed in older adults strength training, the variability of motor output
(Kamen & Knight 2004). This may be because maximal (referred to as steadiness) is reduced during isometric
discharge rates are lower in the untrained state (Kamen (Keen et al. 1994) and slow lengthening contractions
et al. 1995), reflecting a greater potential for change in (Laidlaw et al. 1999). Although multiple factors are
the elderly. However, there is a poor association likely to be responsible (Taylor et al. 2003), one
between increases in muscle strength and maximal candidate mechanism for improved steady motor
motor unit discharge rates after training (r2 = 0.15 for performance after training is a reduction in the
studies in Table 1), presumably because it is only variability of motor unit discharge rate (Moritz et al.
possible to sample from a small proportion of the 2005, Tracy et al. 2005). This hypothesis was exam-
active motor units, and several muscles are likely to ined following a training intervention (Kornatz et al.
contribute to the mechanical action at the relevant joint. 2005) that involved 2 weeks of practicing a steadiness
Nonetheless, these findings suggest that other factors task with a light load (10% of maximum) followed by
4 weeks of strength training with a heavy load (70% of

 2011 The Authors


Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x 121
122
Table 1 Summary of studies examining the effect of strength training on motor unit discharge rate during maximal contractions

Recording Training Training


Study Muscle Subject details technique duration frequency Training intensity Strength gain Discharge rate

Christie & Kamen Tibialis anterior 30 young 4-wire needle 2 weeks 3· week)1 3 sets of 10 isometric › 17% young › 7% young
Neural adaptations to strength training

(2010) 30 old (15 controls/ (quadrifillar) MVCs › 20% old › 24% old
group)
Kamen & Knight Vastus lateralis 8 young 4-wire needle 6 weeks 3· week)1 3 sets of 10 dynamic › 29% young › 15% young
(2004) 7 old (quadrifillar) contractions (85% › 36% old › 49% old
1RM)
Patten et al. Abductor digiti 6 young 4-wire needle 6 weeks 5· week)1 2 sets of 10 isometric › 25% young › 11% young
Æ T J Carroll et al.

(2001) minimi 6 old (quadrifillar) MVCs › 33% old › 23% old


Pucci et al. Vastus lateralis 10 training Tungsten 3 weeks 3· week)1 3 sets of 10 isometric › 35% training › 3% training
(2006) 10 controls microelectrode MVCs group (n.s.)
(All young males)
Van Cutsem et al. Tibialis anterior 5 training Fine-wire needle 12 weeks 5· week)1 10 sets of 10 rapid › 30% training › 29% training*
(1998) 5 controls electrode shortening
(all young subjects) contractions
(30–40% 1RM)
MVC, maximal voluntary contraction; 1RM, one-repetition maximum; n.s., non-significant difference after training.
*This value was calculated without the presence of double discharges (<5 ms).

 2011 The Authors


Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x
Acta Physiol 2011, 202, 119–140
Acta Physiol 2011, 202, 119–140 T J Carroll et al. Æ Neural adaptations to strength training
maximum). After the first 2 weeks of practice, there (De Luca et al. 1982). As a coherence analysis is able to
were parallel declines in motor unit discharge rate detect these multiple features of correlated motor unit
variability in the first dorsal interosseous muscle and activity over a range of frequencies, it is now considered
the steadiness of slow shortening and lengthening the most sensitive approach when examining the
contractions performed by the index finger. No further discharge patterns from a population of active motor
changes in motor unit discharge rate variability or units.
steadiness were evident after 4 weeks of strength One of the most appealing mechanisms for an
training, suggesting that practice of the task was increase in strength with training is an increase in
largely responsible for the changes in motor unit correlated motor unit activity. This has largely been
discharge rate variability and improvements in motor driven by an original investigation from Milner-Brown
performance. Similarly, 6-weeks of strength training et al. (1975), which found a 128% increase in motor
the knee extensor muscles resulted in an increase in unit synchronization after 6 weeks of isometric strength
mean motor unit discharge rates and a decrease in training that involved the index finger. However, this
motor unit discharge rate variability during submaxi- study used a surface EMG based population index of
mal contractions (Vila-Cha et al. 2010). Furthermore, motor unit synchronization, which has been shown to
Knight & Kamen (2004) examined the modulation of be an unreliable estimate of correlated motor unit
motor unit discharge rate during isometric force activity (Yue et al. 1995). Nonetheless, this seminal
tracking tasks requiring skill acquisition. The target study by Milner-Brown et al. (1975) has been widely
force comprised an average of 20% maximal voluntary cited as one of only a few examples of a chronic neural
contraction (MVC) that was modulated by 2% MVC adaptation induced by strength training. Using the more
with sine waves of 0.15 and 0.5 Hz. Throughout the robust cross-correlation procedure between pairs of
performance of 15 trials, they found that an improved concurrently active motor units (see Farmer et al.
precision of force matching was accompanied by 1997), several studies have shown that habitual physical
decreased motor unit discharge rate modulation at activity patterns may influence motor unit synchroni-
the higher of the two modulating frequencies. Taken zation. For example, motor unit synchronization is
together, these studies suggest that training to increase greatest in strength-trained subjects (weightlifters),
maximal strength may have implications for motor unit intermediate in untrained control subjects, and least in
activity during submaximal tasks that may improve skill-trained musicians (Semmler & Nordstrom 1998).
skilled motor performance. More recently, motor unit synchronization has been
shown to be greater in strength-trained males, but only
for the biceps brachii muscle at high forces (80% MVC)
Correlated motor unit activity
and not for the first dorsal interosseous muscle at low
The correlated activity of motor units is now regarded (30% MVC) or high (80% MVC) force levels (Fling
as an important physiological principle for the neural et al. 2009). However, these cross-sectional studies are
control of movement. Computer simulation studies difficult to interpret because no training intervention
have shown that correlated motor unit activity can was performed, and therefore they do not account for
have a substantial influence on EMG and force fluctu- other genetic or lifestyle factors that may contribute to
ations (Yao et al. 2000, Zhou & Rymer 2004), and alterations in motor unit synchronization. In contrast, a
experimental studies suggest that it can vary depending 4-week training intervention that resulted in a 54%
on the details of the task that is performed (see Semmler increase in index finger strength showed no change in
et al. 2002). The correlated discharge of action poten- motor unit synchronization in the first dorsal interos-
tials by motoneurons is caused by common branched seous muscle (Kidgell et al. 2006). This definitive
pre-synaptic inputs or independent inputs that are interventional study using cross-correlation from pairs
themselves synchronized. The effect on motor unit of concurrently active motor units indicates that corre-
activity can be quantified in the time domain as motor lated motor unit activity is unlikely to be important for
unit synchronization or in the frequency domain as the expression of muscle strength with training.
motor unit coherence (Farmer et al. 1997). There is a An alternative explanation for differences in the
strong association between motor unit synchronization strength of correlated motor unit activity in skill and
and high-frequency coherence (10–30 Hz), indicating strength-trained individuals (Semmler & Nordstrom
that similar mechanisms contribute to these two fea- 1998, Semmler et al. 2004) might be that practice of a
tures of correlated motor unit activity (Farmer et al. task results in reduced motor unit synchronization,
1993, Semmler et al. 2004). Furthermore, low-fre- which may be beneficial in performing fine motor tasks
quency (1–2 Hz) motor unit coherence is equivalent at low force levels. This hypothesis was addressed in a
to the common fluctuations in mean motor unit recent study which examined motor unit synchroniza-
discharge rates, which is referred to as common drive tion in hand muscles during a pinching task following

 2011 The Authors


Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x 123
Neural adaptations to strength training Æ T J Carroll et al. Acta Physiol 2011, 202, 119–140

training with light loads to improve force steadiness in single motor unit level has not accelerated markedly
older adults (Griffin et al. 2009). After 4 weeks of light- over the last 10 years. This can largely be attributed to
load isometric training (2 and 4% MVC), they found no technical difficulties associated with recording single
change in motor unit synchronization both within and motor units, particularly at high forces. This limitation
between hand muscles despite significant improvements has been detrimental to studies specifically designed to
in isometric steadiness at these low force levels. In increase strength, as it is the behaviour of high threshold
contrast, there were significant declines in motor unit motor units that are most likely to contribute to the
discharge rate variability in the first dorsal interosseous strength gains, but are least amenable to study with
muscle with training. Thus, it appears that improve- invasive motor unit recording techniques. The contin-
ments in motor performance with skill training or ued development of hardware and software for use with
practice may occur through altered motor unit dis- surface electrode arrays offers many promising avenues
charge rate variability, rather than a change in corre- for advancing knowledge in this field, as they provide a
lated motor unit activity. means of obtaining a high motor unit yield, even during
Finally, evidence from lengthening (eccentric) muscle strong contractions (Farina et al. 2010, Nawab et al.
contractions has provided new insight into the task- 2010). Furthermore, these techniques may also shed
related adjustments of correlated motor unit activity. new light on the coordination of multiple muscle
Lengthening contractions are performed regularly in synergists and antagonists, by facilitating comparisons
everyday lives, and are important considerations for of common drive between muscles and how this may
training because of their potential to produce large contribute to changes in muscle strength with training.
forces with low metabolic cost. Several studies have Finally, for practical reasons, many training studies
shown altered motor unit behaviour during lengthening have been restricted to isometric contractions, but the
contractions (Del Valle & Thomas 2005, Pasquet et al. involvement of lengthening contractions may enhance
2006), supporting the proposal that lengthening con- the changes with training and offer further insight into
tractions require unique neural activation strategies the proposed neural adaptations.
(Enoka 1996). Interestingly, the modulation of corre-
lated motor unit activity also appears to be an impor-
Spinal reflex studies
tant feature of lengthening contractions. For example,
cross-correlation analysis from the same pairs of con-
Technical considerations
currently active motor units in the first dorsal interos-
seous muscle has shown an increase in motor unit Reflex responses produced by electrical stimulation of
synchronization during lengthening, compared with peripheral nerves have the potential to contribute
postural or shortening muscle contractions (Semmler information regarding the sites and mechanisms of
et al. 2002). This increased motor unit synchronization neural adaptation to strength training. The two reflexes
during lengthening contractions may act as a protective that have been applied most commonly in strength
mechanism by distributing the load over a greater training studies are the H-reflex (Magladery & McDo-
number of motor units to reduce the extent of muscle ugal 1950) and the V-wave (Upton et al. 1971), both of
damage that can occur after these types of contractions which rely to a large extent on the monosynaptic circuit
(see McHugh 2003 for review). This suggestion is from Ia afferent fibres to motoneurons. Both reflexes are
supported by recent work showing that motor unit produced by electrical stimulation of a mixed nerve (i.e.
synchronization is increased in biceps brachii motor containing axons of motoneurons and afferents of
units for at least 24 h after the muscle has been various origin). The afferent volley generated ultimately
damaged with repetitive lengthening contractions causes a relatively synchronous discharge of motoneu-
(Dartnall et al. 2008). These data suggest that length- rons, and the earliest part of the response is probably
ening muscle contractions, including the involvement of due purely to the monosynaptic connection between Ia
naturally occurring movements such as the stretch- afferents and motoneurons (Magladery et al. 1951).
shortening cycle (see Nicol et al. 2006), may be impor- With H-reflexes, the response initially grows with
tant in the neural adaptation and the modulation of stimulation intensity as the afferent volley increases in
correlated motor unit activity with training. size, but then progressively falls to zero as the intensity
approaches levels supramaximal for activating motor
axons (i.e. to produce a maximal M-wave). It is
Moving forward with human motor units
generally accepted that this is because of collision
Over the last 3 decades, evidence has accumulated to between orthodromic action potentials produced by the
suggest that certain types of strength or resistance reflex with antidromic action potentials generated
training can alter the discharge properties of human directly in the motor axons by the electrical stimulus.
motor units. However, the progress in this field at the In contrast, the V-wave is produced by supramaximal

 2011 The Authors


124 Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x
Acta Physiol 2011, 202, 119–140 T J Carroll et al. Æ Neural adaptations to strength training
stimulation during voluntary contraction (typically non-monosynaptic contributions to the reflex (e.g.
MVC). The cancellation of action potentials that Burke et al. 1984, Marchand-Pauvert et al. 2002, see
abolishes the H-reflex at high stimulus intensities is Burke & Gandevia 1999 for review). While the earliest
prevented in some motoneurons by action potentials part of the response is via monosynaptic Ia inputs, this
produced via volition. That is, in some motoneurons, a initial excitation is curtailed by disynaptic inhibition
voluntary action potential will collide with the (i.e. produced via Ia or Ib afferent activation of an
antidromic action potential produced by the stimulus, inhibitory interneuron; Marchand-Pauvert et al. 2002).
and ‘clear’ the pathway for subsequent reflex activation. Thus, as pointed out by Burke & Gandevia (1999), the
The number of motoneurons that are free to respond to size of the H-reflex is determined by the balance
the afferent volley will depend on the number and firing between monosynaptic Ia excitation and disynaptic
rate of the motoneurons involved in the voluntary inhibition. A change in H-reflex or V-wave size could be
contraction, and the conduction time from the stimu- brought about by an alteration in transmission efficacy
lation site to the spinal cord. The ultimate size of the in either of these circuits. Clearly, as pointed out
reflex will also depend on motoneuron responsiveness previously by many authors (e.g. Burke & Gandevia
and the efficacy of synaptic transmission between 1999, Zehr 2002, Misiaszek 2003), the H-reflex does
afferents and motoneurons. Although these two reflexes not provide a measure of ‘motoneuron excitability’.
can provide information regarding the efficacy of the Ia Finally, if one considers only monosynaptic Ia con-
afferent reflex pathway and the output from the tributions, changes in the size of the H-reflex could be
motoneuron pool during voluntary contraction, there because of changes in the responsiveness of the moto-
are a number of technical issues that complicate their neurons or alterations in the synaptic transmission
interpretation. Because many of the complications with efficacy at the Ia afferent terminals (see Zehr 2002,
H-reflex interpretation are often neglected, some of the Misiaszek 2003 for reviews). With respect to moto-
key points will be summarized in the following sections. neuron responsiveness, intrinsic membrane properties,
The reader is referred to Burke & Gandevia (1999) for firing rate and the effects of neuromodulators such as
detailed coverage. serotonin could contribute (Capaday 1997, Matthews
The first complication in interpreting changes in 1999). Alterations to the synaptic transmission efficacy
H-reflex amplitude because of an intervention such as can occur because of pre-synaptic inhibition of Ia
strength training is the difficulty in ensuring an identical afferent terminals and homosynaptic post-activation
size of the afferent volley elicited by the stimulation depression (see Hultborn & Nielsen 1998, Pierrot-
pulse. One way of attempting to control this is by Deseilligny & Burke 2005 for extensive review).
matching the amplitude of a small M-wave between
conditions, and thereby ensuring a similar number of
H-reflex and V-wave studies
motoneurons are excited by the stimulus (e.g. Zehr
2002, Misiaszek 2003). However, this procedure does Since the earliest cross-sectional (Milner-Brown et al.
not account for potential changes in axonal excitability 1975), and longitudinal studies (Sale et al. 1982, 1983),
(e.g. Bostock & Grafe 1985). This is a critical consid- there have been many reports of H-reflex and V-wave
eration because axonal excitability is susceptible to responses to strength training (summarized in Table 2).
change in ways that differ for large diameter afferents With the exception of the earliest studies (Sale et al.
versus motoneurons (e.g. Mogyoros et al. 1996, Vagg 1982, 1983), all of these investigations have involved
et al. 1998). If strength training were to alter axonal muscles that act at the ankle joint. There have been no
excitability of afferents or motoneurons, this could reports of an increase in H-reflex amplitude at rest as a
produce a change in H-reflex amplitude in the absence consequence of strength training (Aagaard et al. 2002,
of central effects [i.e. effects within the central nervous Scaglioni et al. 2002, Del Balso & Cafarelli 2007,
system (CNS)]. However, measures of H-reflex Holtermann et al. 2007, Duclay et al. 2008, Schubert
responses provided by H-reflex recruitment curves et al. 2008, Fimland et al. 2009a, Ekblom 2010) or
(i.e. recording the reflex size at a range of stimulus tetanic electrical stimulation of muscles (Maffiuletti
strengths from below threshold to supramaximal levels) et al. 2003, Gondin et al. 2006). In contrast, strength
should be less susceptible to this potential confound. training was reported to increase the amplitude of
The issue is also less likely to be of concern for V-waves V-waves in all studies involving plantar flexors (Sale
provided that the stimulus intensity was sufficiently et al. 1983, Aagaard et al. 2002, Gondin et al. 2006,
supramaximal to account for activity dependent reduc- Del Balso & Cafarelli 2007, Holtermann et al. 2007,
tions in axonal excitability (see Burke & Gandevia 1999 Duclay et al. 2008, Fimland et al. 2009a,b, Ekblom
for details). 2010) but not for the thumb abductors (Sale et al.
The second complication for interpreting the size of 1982). Thus, for the Soleus muscle, there is clear
H-reflexes is that there are almost certainly considerable evidence that resting H-reflexes are unaffected, and

 2011 The Authors


Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x 125
Table 2 Summary of studies examining the effect of strength training on H-reflex and V-wave size

126
Training Training
Study Muscle group Subject details Methods schedule intensity Strength gain Results

Sale et al. (1982) Thumb abductors 11 trained V-wave, MVC 18 weeks Dynamic › 41% MVC M V-wave
74 sessions 5–6 sets up to 1RM
Isometric
2 sets 10 MVC
Sale et al. (1983) Plantar flexors 14 trained V-Wave, MVC 9–21 weeks Dynamic Not reported › 50% V-wave. Pooled
toe extensors F-wave 5–6 sets up to 1RM across muscles
elbow flexors Isometric M F-wave
finger flexors 2 sets 10 MVC
Aagaard et al. Plantar flexors 14 trained H-reflex, rest and 90% 14 weeks Dynamic › 20% MVC M H-reflex at rest
Neural adaptations to strength training

(2002) MVC 38 sessions 3–10RM › 23% concentric › 55% V-wave,


V-Wave, 90% MVC multiple exercises 4–5 › 19% H-reflex 90%
sets each MVC
Scaglioni et al. Plantar flexors 14 trained H-reflex, rest 16 weeks Dynamic › 18% MVC M H-reflex
(2002) (65–85 years) 48 sessions 50–80% 1RM › 24% 1RM
Æ T J Carroll et al.

1 set of 10
Maffiuletti et al. Plantar flexors 8 trained H-reflex, rest 4 weeks 45 electrical Not reported M H-reflex
(2003) 6 control 16 sessions stimulations (max
tolerable)
50–70% MVC
Lagerquist et al. Plantar flexors 10 trained H-reflex, 10% MVC 5 weeks Isometric › 15% MVC › H-reflex during 10%
(2006) 6 control 15 sessions MVC MVC
5 sets of 8
Gondin et al. Plantar flexors 12 trained H-reflex, rest and MVC 5 weeks 40 electrical › 22% MVC M H-reflex
(2006) 7 Control V-Wave, MVC 15 sessions stimulations › 80% V-wave
(max. tolerable)
60% MVC
Del Balso & Cafar- Plantar flexors 10 trained H-reflex, rest, 10% 4 weeks Isometric › 20% MVC M H-reflex
elli (2007) 10 control MVC 12 sessions 6 sets of 10 MVC › 54% V-wave
V-Wave, 50, 75, 100%
MVC
Beck et al. (2007) Dorsiflexors, 8 trained H reflex during fast 4 weeks Dynamic ballistic Not reported M H-reflex
Plantar flexors 10 control contraction 16 sessions 30–40% 1RM
4–6 sets of 10
Holtermann et al. Plantar flexors 12 trained H-reflex at rest and 3 weeks Isometric › 18% MVC M H-reflex at rest
(2007) 12 control during 20 and 60% 9 sessions MVC › H-reflex, 20, 60%
MVC 5 sets of 10 MVC

 2011 The Authors


Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x
Acta Physiol 2011, 202, 119–140
Table 2 Continued

Training Training

 2011 The Authors


Study Muscle group Subject details Methods schedule intensity Strength gain Results

Duclay et al. (2008) Plantar flexors 10 trained H-reflex at rest & 7 weeks Dynamic eccentric, › 22% MVC M H-reflex at rest
8 control during passive 18 sessions 120% 1RM › 31% max › V-wave in MVC and
Acta Physiol 2011, 202, 119–140

movement 6 sets of 6 eccentric max eccentric


H-reflex and V-wave › 28% max › H-reflex in MVC, max
during max isometric, concentric eccentric and concentric
concentric and
eccentric contractions
Schubert et al. Plantar flexors 12 trained H-reflex and 4 weeks Dynamic ballistic › RFD M H-reflex
(2008) 11 control TMS-conditioning of 16 sessions 30–40% 1RM MVC not reported › conditioned H-reflex
H reflex during rest and 4–6 sets of 10 size in ballistic task but
ballistic tasks not rest
Geertsen et al. Dorsiflexors 12 trained Disynaptic reciprocal 4 weeks Isometric › 20% MVC › reciprocal inhibition
(2008) 6 control inhibition of plantar 12 sessions MVC ballistic
flexors at onset ballistic 3 sets of 16
dorsiflexion
Fimland et al. Plantar flexors 10 trained H-reflex at rest 8 weeks Dynamic ballistic › 20% MVC › 50% V-wave in MVC
(2009a) 9 control H-reflex and V-wave 24 sessions multi-joint extension › 44% 1RM leg M H-reflexes
during MVC 85–90% 1RM press
T J Carroll et al.

4 sets of 4
Fimland et al. Plantar flexors 15 trained H-reflex and V-wave 4 weeks Isometric › 44% MVC › 72% V-wave in MVC
(2009b) 11 control during MVC 16 sessions MVC M H-reflexes
6 sets of 6
Ekblom (2010) Plantar flexors 9 trained H-reflex and V-wave 5 weeks Dynamic › 19% mean M H-reflex
11 control during max concentric 15 sessions eccentric focused for concentric/eccentric › 77% V-wave during

Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x


and eccentric PF maximal eccentric and
H-reflex at rest 5–6 sets of 5RM concentric actions

max, maximal; MVC, maximal voluntary contraction; RM, repetition maximum.

127
Æ Neural adaptations to strength training
Neural adaptations to strength training Æ T J Carroll et al. Acta Physiol 2011, 202, 119–140

V-wave amplitudes are enhanced, by strength training. sure of an identical reflex excitation of motoneurons),
The results for H-reflexes elicited during voluntary then there is no basis for determining the origin of the
contractions are less consistent. There have been reports extra motoneuronal drive. For example, spinal reflex
both that strength training increases H-reflex size pathways are known to contribute considerably to the
(Aagaard et al. 2002, Lagerquist et al. 2006, Holter- activation of motoneurons during voluntary contrac-
mann et al. 2007, Duclay et al. 2008) and does not tion (e.g. Macefield et al. 1993).
change H-reflex size (Gondin et al. 2006, Beck et al.
2007, Del Balso & Cafarelli 2007, Fimland et al.
Conditioning-test reflex approaches
2009a,b, Ekblom 2010), when evoked during contrac-
tions ranging in strength from 10 to 100% of MVC. Given the difficulty in identifying the specific spinal
The discrepancies could be due partly to differences in circuits responsible for any changes that might occur in
the details of how the reflexes were elicited and H-reflex (or V-wave) size, it is worth considering the
measured. For example, some studies determined the potential of related methods that might provide more
maximal H-reflex amplitude, whereas others assessed specific information about the nature of neural adap-
H-reflex size at matched M-wave amplitudes, or calcu- tation to strength training. By careful application of
lated the slope of the stimulus intensity versus response conditioning stimulation to various homonymous and
curve. Each of these methods might be differentially heteronymous muscle afferents or skin afferents, it is
affected by methodological factors that can influence possible to draw inferences about the efficacy of
H-reflex amplitude such as axonal excitability. specific spinal reflex pathways (see Pierrot-Deseilligny
When a change in H-reflex was detected, the result & Burke 2005 for extensive review). For example,
was taken as evidence of neural adaptation to strength Geertsen et al. (2008) used conditioning stimulation of
training mediated at a spinal level. Whether this is the peroneal nerve to suppress the soleus H-reflex in
entirely accurate depends on one’s perspective on order to study the effect of explosive strength training
defining a ‘site’ of adaptation. While it is accepted on disynaptic reciprocal inhibition (i.e. via Ia inhibi-
that the H-reflex is subject to modulation within the tory interneurons that reduce coactivation of antago-
spinal cord, it is quite possible that changes in nist muscles). They found that 4 weeks of ballistic
supraspinal inputs to the spinal circuitry could increase strength training for the dorsiflexors increased disy-
H-reflex size (i.e. by modulating pre-synaptic inhibition naptic reciprocal inhibition from peroneal afferents to
or the responsiveness of inhibitory interneurons). the antagonist soleus when assessed at the onset of
Increases in V-wave amplitude have frequently been ballistic dorsiflexion contraction, but not at rest. The
cited as evidence of increased motoneuron activation. lack of change in disynaptic reciprocal inhibition at
There has also been a strong tendency to ascribe the rest implies that there was no change in the default
putative increased motoneuron activation to adapta- properties of the pathway per se, and that the
tions at supraspinal levels, particularly when V-wave increased inhibition prior to ballistic contraction might
changes were observed in parallel with H-reflexes (e.g. reflect a training-induced descending modulation of
Aagaard et al. 2002, Del Balso & Cafarelli 2007). spinal interneuronal responsiveness that is tailored to
While an increased supraspinal drive to motoneurons is the functional requirements of the training task (i.e. to
certainly a plausible explanation for increased V-wave ensure that antagonist muscles do not contribute
amplitudes, there are numerous alternative possibilities plantar flexion torque at the onset of intended dorsi-
that cannot be discounted. More generally, the notion flexion).
that comparisons between V-wave and H-reflex allow
dissociation between spinal and supraspinal adapta-
Moving forward with reflex studies
tions is ill-founded. For example, the different stimulus
intensities associated with the two methods will acti- Despite the many complications with interpreting
vate of different populations of afferents and moto- H-reflex and V-wave results, it appears that for the
neurons (i.e. antidromically), which might also result in plantar flexors, resting H-reflexes are unaffected by
different oligosynaptic influences on the reflex ampli- strength training and V-waves during MVC are
tude (e.g. greater antidromic activation of motoneurons increased. The physiological implications of these
produced by supramaximal stimulation to generate the results are not definitive, but it is likely that there is
V-wave might activate more Renshaw cells), and the little ‘baseline’ alteration to the Ia-afferent-motoneu-
reflexes might be differentially influenced by activity ronal circuit, and an increase in motoneuronal output
dependent changes in axonal excitability. Furthermore, during MVC. An understanding of spinal responses to
even if an increase in V-wave amplitude could be strength training might be advanced by the applica-
ascribed entirely to an increase in the number of tion of conditioning-test approaches to study modu-
voluntarily recruited motoneurons (i.e. if one could be lation in specific reflex circuits. While these

 2011 The Authors


128 Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x
Acta Physiol 2011, 202, 119–140 T J Carroll et al. Æ Neural adaptations to strength training
approaches are powerful in the sense that they have size of D-waves produced some distance from the
potential to provide specific information about par- corticospinal cell soma is little influenced by the state of
ticular afferent and interneuronal pathways, it is the cortical network (see Di Lazzaro et al. 2004 for
important to note that they provide only indirect review), but the size of I-waves is determined by the net
evidence (i.e. inferences obtained require confirmation result of multiple synaptic inputs to the corticospinal
from more than one method) and can be technically cells. These inputs could arise both from local cortical
challenging. interneurons and projections from distant cortical and
subcortical sites, are transmitted from the site of (TMS)
activation to the corticospinal cells via mono- and
TMS studies
oligosynaptic circuits, and contain both excitatory and
inhibitory elements (see Reis et al. 2008 for recent
Technical considerations
review). Thus, the net descending volley produced in the
Transcranial magnetic stimulation can painlessly acti- corticospinal tract by a TMS pulse is potentially
vate neurons in the human cerebral cortex through the influenced by the intrinsic responsiveness of the corti-
scalp (Barker et al. 1985), and has been used to study cospinal cells (i.e. which is influenced by firing rate and
neural transmission from the motor cortex to the membrane properties), the synaptic efficacy of local and
muscles in many behavioural contexts. An index of distant connections onto corticospinal cells and local
the responsiveness of the entire pathway from brain to interneurons, and the axonal excitability of the cells
muscle (often termed ‘corticospinal excitability’) can be initially responsive to the TMS pulse. Given the
obtained from the size of the compound muscle action complexity of the system between the sites of stimula-
potentials (CMAP) recorded at the muscle via electro- tion and measurement of the response, the conception
myography. These responses are known as motor that alterations in ‘cortical excitability’, as suggested by
evoked potentials (MEPs). Given that MEPs are gener- TMS studies, might provide a mechanism of strength
ated in response to stimulation of the brain, it can be gains produced by training represents a somewhat
tempting to interpret alterations in MEP size as being superficial consideration of the physiological changes
caused by a change in the properties of the motor that might be involved.
cortex. This is not appropriate, however, as the size of In addition to the anatomical organization and
TMS responses is subject to potential modulation at functional state of cortical circuits (which modulate
each point in the pathway from brain to muscle. Thus, the descending volley), the size of the MEP induced by
in order to properly interpret the effects of an interven- TMS is also influenced by factors within the spinal cord.
tion on MEP size, consideration of all of the biophysical The fact that there are numerous monosynaptic corti-
and physiological factors that determine muscle cospinal connections to distal limb motoneurons in
responses to TMS is required. This is particularly primates (see Porter & Lemon 1993 for review), has
important for longitudinal studies involving strength perhaps led to excessive focus on the corticospinal cell –
training where there might be substantial changes in motoneuronal connection as a potential site of MEP
spinal cord circuitry (see Spinal reflex studies above). modulation. It is correct that the responsiveness of
Transcranial magnetic stimulation can activate cort- motoneurons (i.e. influenced by firing rate and
icospinal cells (i.e. output cells from the motor cortex membrane properties) and the efficacy of the cortico-
that project to the spinal cord) both directly and spinal-motoneuronal synapse can affect the MEP size.
indirectly, depending on the magnitude and orientation However, because many corticospinal cells terminate on
of the induced electric field to the cortical network (i.e. interneurons in the spinal cord (see Porter & Lemon
which are influenced by coil size, shape and orientation, 1993 for review), the net motoneuronal output pro-
stimulus intensity, and the anatomy of the skull and duced by a TMS pulse can be affected by the synaptic
cortex; e.g. Barker 1991, Ruohonen & Ilmoniemi efficacy and intrinsic responsiveness of the components
1999). Direct activation of corticospinal cells results in of spinal interneuronal circuits. It is quite possible that
a descending corticospinal volley termed a ‘D-wave’, MEP size might change despite an identical output from
that can be recorded by electrodes implanted in the the cortex (i.e. due entirely to spinal factors), and this is
cervical epidural space (e.g. Di Lazzaro et al. 1998a), true even if the level of background EMG does not
and inferred from peri-stimulus time histograms from change. Comparisons of MEP size with H-reflex size,
single motor unit recordings (e.g. Day et al. 1989). often erroneously interpreted as an index of ‘motoneu-
Indirect activation of corticospinal cells results in one or ronal excitability’ (see Spinal reflex studies section), do
more ‘I-waves’ (n.b. numbered in order of appearance not allow dissociation of spinal from cortical effects
latency), the earliest of which (i.e. the I1 wave) appears because the afferent and descending volleys might
approximately 1.5–2 ms after the D-wave (Day et al. activate different interneuronal circuits, and because
1989, Di Lazzaro et al. 2004, Brocke et al. 2005). The the monosynaptic components of the responses traverse

 2011 The Authors


Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x 129
Neural adaptations to strength training Æ T J Carroll et al. Acta Physiol 2011, 202, 119–140

different synapses (i.e. the corticospinal-motoneuronal M-wave in most studies, the possibility that the results
synapse versus the Ia-afferent-motoneuronal synapse; could be because of muscular factors alone can gener-
both of which are subject to modulation with voluntary ally be excluded). However, it is important to consider
muscle activation; Hultborn & Nielsen 1998, Petersen carefully what conditions represent ‘rest’. MEPs are
et al. 2003). Only the response to a descending volley typically considered to be elicited at rest if the target
that travels in the same corticospinal axons as those muscle exhibits EMG silence. However, even in the
activated by TMS, such as transmastoid electrical or absence of overt motoneuron activity, contextual fac-
magnetic stimulation, can provide unambiguous tors such as muscle history (Stuart et al. 2002) and
evidence about the spinal or cortical location of MEP motor imagery (e.g. Hashimoto & Rothwell 1999) can
modulation. influence MEP size. As it is difficult to control all
Finally, since the MEP is recorded from the muscles relevant (i.e. especially cognitive) factors, there is likely
via electromyography, its size is subject to all of the to be some ‘noise’ in the assessment of corticospinal
factors that can affect the amplitude of a CMAP. These responsiveness at rest. If strength training induces
include electrode impedance and the location of elec- corticospinal adaptations that are subtle, resting MEPs
trodes relative to the muscle architecture and innerva- may lack the precision to detect them.
tion zones (which can be influenced, for example, by The variability of the results across studies is even
muscle length and contractile status; Lee & Carroll more apparent for assessments of MEP size during
2005, Frigon et al. 2007), the characteristics of tissue muscle activation. The majority of studies have assessed
between the muscle and the skin, the number and MEP size during weak isometric contraction of the
functional state of ion channels in the muscle mem- trained muscles. Under these conditions MEPs have
brane, the relative timing of action potential firing in been shown to increase in the plantar flexor muscles
different muscle fibres, and synaptic transmission at the (Griffin & Cafarelli 2007), be unchanged in finger
neuromuscular junction. Thus, it is conceivable that muscles (Carroll et al. 2002, Kidgell & Pearce 2010) or
MEP size might change despite an identical motoneu- trend lower in arm muscles (Jensen et al. 2005, Carroll
ronal output in response to a TMS pulse (i.e. due et al. 2009). Changes in MEP size during contractions
entirely to peripheral factors). In order to properly of matched strength could potentially be mediated by all
interpret MEP size, it is essential to normalize the of the factors outlined for resting responses, but are
responses to a standardized CMAP such as the maximal likely to be affected to a relatively greater extent by
M-wave (i.e. which is elicited by supramaximal stimu- motoneuron firing rate and intrinsic motoneuron firing
lation of the motor nerve). properties (Matthews 1999). Only Carroll et al. (2002,
2009) considered MEP amplitudes during strong iso-
metric contractions, and reported reductions in MEP
MEP studies
size for contraction strengths over about 50% MVC in
The details of studies that have examined the effect of both finger and wrist muscles. These reductions in
strength training protocols on EMG responses to TMS corticospinal responsiveness were likely mediated
are summarized in Table 3. These studies show that largely at the spinal level, as similar results were
there is no clear pattern of change in MEP size after obtained when the descending tracts were activated by
short-term strength training (i.e. 4 weeks). It is pos- transcranial electric stimulation (Carroll et al. 2002)
sible that the inconsistent results might be because of and stimulation of the corticospinal tract via stimula-
variations in testing and training protocols, and in the tion at the cevicomedullary junction (Carroll et al.
muscle groups studied. For example, the size of MEPs 2009). Although both of these techniques are less
recorded from muscles at rest was unchanged after affected by the state of cortical circuits than TMS (see
training of finger (Carroll et al. 2002, Hortobagyi et al. Di Lazzaro et al. 1998a,b for evidence in relation to
2009), wrist (Carroll et al. 2009) or leg muscles (Griffin transcranial electrical stimulation), stimulation at the
& Cafarelli 2007, Schubert et al. 2008), but was cervicomedullary junction is the superior method of
decreased by training of proximal arm muscles (Jensen dissociating cortical from subcortical effects (Taylor &
et al. 2005). Resting MEP size has been used to provide Gandevia 2004). This is because corticospinal axons are
an index of ‘baseline’ corticospinal responsiveness that activated well caudal to the cortical circuits (excluding
represents the default characteristics of the pathway direct influence of cortical excitability to the responses),
from motor cortex to muscle. From this perspective, and peripheral responses to both TMS and cervicome-
changes in resting MEP size might reflect training- dullary stimulation arise from at least some identical
induced adaptations in the efficacy of synapses and corticospinal axons (Ugawa et al. 1991, Gandevia et al.
axonal and/or cellular excitability at cortical sites, or in 1999, Taylor et al. 2002). Unfortunately, the technique
synaptic efficacy or cellular excitability at subcortical is painful (particularly if responses at rest are sought),
sites (n.b. as responses are normalized to the maximal and it will probably be difficult to recruit large samples

 2011 The Authors


130 Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x
Table 3 Summary of studies examining the effect of strength training on MEP size

Muscle Subject Training


Study group details Methods schedule Training intensity Strength gain Results

 2011 The Authors


Carroll et al. 2002 Finger abductors 8 trained TMS at rest 4 weeks Dynamic › 30% MVC M Resting MEPs
8 control TMS and TES during 12 sessions 70–85% initial 1RM fl TES and TMS MEPs
0–60% MVC 4 sets of 8 at >50% MVC
Jensen et al. 2005 Elbow flexors 8 trained TMS at rest 4 weeks Dynamic › 31% 1RM fl Resting MEP max size
Acta Physiol 2011, 202, 119–140

8 control TMS during 10% MVC 12 sessions 10–6RM › 12% MVC and slope
5 sets of 10–6 n.s. trend for same at
10% MVC
Griffin & Cafarelli Dorsiflexors 10 trained TMS at rest 4 weeks Isometric › 18% MVC › MEP size at 10%
2007 10 control TMS during 10% MVC 13 sessions MVC MVC
6 sets of 10 M at rest
Beck et al. 2007 Dorsiflexors 8 trained TMS and H reflex during 4 weeks Ballistic, dynamic Not reported M H-reflex
plantar flexors 10 control fast contraction 16 sessions 30–40% 1RM M SICI or ICF
SICI and ICF at rest 4–6 sets of 10 › TMS during
dorsiflexion in TA
Schubert et al. 2008 Plantar flexors 12 trained TMS-conditioning of H 4 weeks Ballistic, dynamic › Rate of force fl conditioned H-reflex
11 control reflex during rest and 16 sessions 30–40% 1RM development, size in ballistic task but
ballistic tasks 4–6 sets of 10 MVC not not rest
reported
Carroll et al. 2009 Wrist abductors 8 trained TMS and CMEPs during 4 weeks Dynamic › 11% MVC fl MEP and CMEP at
9 control 0–75% MVC 12 sessions 70–85% initial 1RM high contraction
T J Carroll et al.

4 sets of 8 strength
M MEP at rest
Hortobagyi et al. Finger abductors 8 trained TMS at rest 4 weeks Isometric › 38% MVC M MEPs
2009 8 control 10 sessions 70–80% MVC
5 sets of 10
Kidgell & Pearce Finger abductors 8 trained TMS during 5 and 20% 4 weeks Isometric › 34% MVC M MEPs

Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x


2010 8 control MVC 12 sessions MVC
6 sets of 10
TMS, transcranial magnetic stimulation; TES, transcranial electrical stimulation; SICI, short-interval intracortical inhibition; ICF, intracortical facilitation; CMEPs, cervico-meduallary evoked
potentials; MVC, maximal voluntary contraction; RM, repetition maximum; n.s., non-significant; TA, tibialis anterior.

131
Æ Neural adaptations to strength training
Neural adaptations to strength training Æ T J Carroll et al. Acta Physiol 2011, 202, 119–140

for studies involving this technique. Carroll et al. (2002) might speculate that the predominant component of the
speculated that the decreases in MEP size after strength neural adaptation to strength training should be asso-
training at high forces could be explained by changes in ciated with refining circuitry to facilitate activation of
the firing rate of the motoneurons or an increase in the synergists and suppress activation of antagonists. This
amplitude or duration of motoneuronal after-hyperpo- reasoning reinforces the notion that MEPs recorded
larisation potential. from agonist muscles may lack the sensitivity to detect
The group of Beck et al. (2007) and Schubert et al. corticospinal adaptations in many contexts. How then
(2008) assessed TMS responses after a form of strength could TMS be used to identify the neural mechanisms
training involving ballistic contractions. The size of that underlie putative changes in coordination between
MEPs was greater after training when elicited during agonist, synergist and antagonist muscles?
the execution of dynamic contractions that resembled
the training task, but not at rest (Beck et al. 2007). The
Use of TMS-induced twitch responses
data indicate a context-specific increase in corticospinal
responsiveness after training. In contrast, the size of One obvious approach to studying possible changes in
H-reflexes conditioned by subthreshold TMS pulses coordination between muscles after strength training is
was reduced in a task-dependent manner (Schubert to measure MEPs in antagonist and synergist muscles.
et al. 2008). The H-reflex conditioning technique However, a major impediment to experiments of this
involves assessing the degree to which a TMS pulse kind would be the difficulty in obtaining selective
(conditioning pulse), that is subthreshold for producing recordings from all of the individual muscles that might
an MEP on its own, increases the size of an H-reflex be relevant. At most joints, there are many muscles that
(test pulse) that is elicited near simultaneously (see can contribute non-negligible moments, and it is pos-
Nielsen et al. 1993). As the precise conditioning-test sible that important functional effects might be brought
interval is specified so that the H-reflex is facilitated about by the combined effect of small changes (i.e. that
only by the earliest (monosynaptic) component of the are difficult to measure) in multiple muscles. Some of
descending volley, the conditioned H-reflex results of these muscles will be in close proximity to one-another
Schubert et al. (2008) are consistent with a decrease in (i.e. making selective recording difficult), and some will
the size of the motor cortical output in response to be deep (i.e. making detection with surface electrodes
TMS. Although adaptations at the cortico-motoneuro- difficult). Another approach is to record the twitch
nal synapse (see Petersen et al. 2003) could also explain torque evoked by TMS, which will reflect the net
a change in conditioned H-reflex amplitude, the lack of response in all agonist, synergist and antagonist mus-
change in this variable at rest argues against this cles. The major difficulty here is to ensure that
possibility. It is difficult to draw general conclusions contractile factors do not contribute to any change in
from the conflicting results; that is, task-specific the responses. Carroll et al. (2009) found that TMS-
increase in corticospinal excitability revealed by MEPs evoked twitches were larger during weak background
versus task-specific decrease in cortical excitability contraction in the wrist muscles after 4 weeks of
revealed by TMS conditioning of H-reflexes. However, strength training involving wrist abduction. Spinal
the studies illustrate the importance of considering task factors were likely dominant, since similar results were
context when attempting to categorize the neural obtained for twitches evoked by cervico-medullary
responses to training. stimulation. There were no differences in twitch ampli-
Given the inconsistent results from studies that have tudes evoked by supramaximal nerve stimulation
focused on the effect of strength training on MEP size in (at rest or during identical voluntary contractions),
muscles directly targeted during training, progress suggesting that muscular factors did not underlie the
toward an understanding of the neural responses to results. Similarly, Lee et al. (2009) reported increased
training might be facilitated by reflecting on putative TMS-evoked twitch size during high force contractions
mechanisms that have been proposed on conceptual in a larger study with an overlapping subject cohort.
grounds. If a change in the CNS is to enhance strength, They also found that the direction of evoked twitches
it must act by increasing the activation of motoneurons shifted toward the training direction. The differences in
that contribute to torque in the desired direction (i.e. by the results between the two studies (i.e. no effect on
recruiting additional motoneurons or increasing the TMS-evoked twitches at high forces in the Carroll et al.
firing rate of motoneurons that innervate agonist and 2009 study), despite identical training parameters, is
synergist muscles), or by reducing the activation of probably because of subtle differences in TMS intensity
motoneurons that oppose torque in the desired direction and the contractile conditions (i.e. whether subjects
(i.e. antagonist muscles). Since the ability of the CNS to matched EMG or torque as a percentage of MVC).
activate many muscles (i.e. when acting as agonists) is Although the results are consistent with the hypothesis
near maximal (see Gandevia 2001 for review), one that strength training causes corticospinal adaptations

 2011 The Authors


132 Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x
Acta Physiol 2011, 202, 119–140 T J Carroll et al. Æ Neural adaptations to strength training
that facilitate coordination between the muscles the trained limb (i.e. produced by isometric training).
involved in training, many questions remain about the These results suggest that the motor cortex is better able
precise mechanisms involved. to drive motoneurons during MVC of the untrained,
homologous muscles after strength training. This could
be due to adaptations at any point upstream of the
Cortical voluntary activation
motor cortical output cells; that is, in the circuitry of the
Although the technique of twitch interpolation (Merton primary motor cortex, or at any of the secondary motor
1954) is not the focus of this review, it is worth noting cortical areas or subcortical areas that shape the output
that a version of the method that involves TMS might of the primary motor cortex. It seems likely that similar
have some benefits over ‘traditional’ nerve stimulation mechanisms might contribute to strength gains in the
methods for identifying changes in the ability of the trained limb, which might only be detectible when
CNS to activate muscles after strength training (see training causes a considerable increase in MVC.
Todd et al. 2003). Twitch interpolation with nerve
stimulation is limited to assessing voluntary activation
Use of repetitive TMS (rTMS)
in a single muscle (with motor point stimulation) or in a
group of muscles innervated by a single peripheral In the studies described above, TMS was used to probe
nerve. The nerve stimulation technique also provides no the responsiveness of the pathway from motor cortex to
information about the site of any failure in the capacity muscle. However, in repetitive mode, TMS can also be
of the CNS to drive the muscles maximally. In contrast, used to transiently disrupt ongoing cortical processing,
if TMS can elicit extra force during an MVC, it or to induce persistent (i.e. that outlast the stimulation
indicates that the output from the motor cortex was by a few minutes to an hour) changes in the respon-
suboptimal for driving the motoneurons (see Gandevia siveness of cortical (n.b. and/or spinal) circuits (see
et al. 1996, Todd et al. 2003). Furthermore, given the Fitzgerald et al. 2006, Ziemann et al. 2008 for recent
relatively poor spatial resolution of TMS, the proximity review). In the only strength training study to date to
of cortical representations of functionally related mus- use rTMS in this way, Hortobagyi et al. (2009) applied
cles, and the fact that the corticospinal responsiveness a low-frequency protocol that is known to induce
of muscles activated voluntarily is considerably higher lasting corticospinal inhibition (1 Hz for 15 min at
than resting muscles, the TMS twitch interpolation 110% of resting motor threshold; Chen et al. 1997) to
method can, in many cases, provide information about the motor cortex after every session in a 4 week
the ability of the motor cortex to drive all of the relevant strength training program. The participants who
agonist and synergist muscles in any given contraction. received rTMS showed impaired strength gains relative
It is important to note that for TMS methods of twitch to control subjects by the end of training, which implies
interpolation to be valid, it is essential that the stimulus that the cortical or spinal circuits activated by the rTMS
produces a near maximal response in the relevant were important substrates for the adaptations that
muscles. Any degree to which the stimulus is submax- increased strength. It is important to note that although
imal will introduce variability into the measurements, rTMS effects are commonly attributed to cortical
and substantially submaximal stimulations would result factors by default, a spinal origin for effects cannot be
in invalid results. Nevertheless, the validity and reli- discounted when the stimulus is above the threshold for
ability of the method has been demonstrated in the wrist a descending volley (n.b. there will be a descending
extensors (Lee et al. 2008), elbow flexors (Todd et al. volley from the brain to the spinal cord at stimulus
2003, 2004) and knee extensors (Sidhu et al. 2009). intensities well below resting motor threshold; e.g.
Lee and colleagues used twitch interpolation with Di Lazzaro et al. 1998b).
TMS to study the influence of strength training on the Although they have yet to be applied to the context of
ability of the motor cortex to drive wrist muscles after strength training per se, rTMS protocols that induce
4 weeks of strength training (Lee et al. 2009). They effects that resemble synaptic plasticity in the motor
found that cortical voluntary activation of the wrist cortex might allow insights into the nature of corti-
abductors was high at baseline (94%), and did not cospinal responses to training. If strength training does
change as a consequence of strength training. However, cause cortical adaptations, it seems likely that short-
the dynamic training was not specific to the isometric term changes might be mediated by synaptic processes
conditions under which voluntary activation was similar to those thought to be involved in motor
assessed, and the improvement in MVC was relatively learning, (e.g. Carroll et al. 2001, Carson 2006, c.f.
small (11%). Lee et al. (2010) subsequently showed Adkins et al. 2006). Long-term potentiation (LTP) is a
that cortical voluntary activation increased for the wrist likely candidate for the cortical reorganization that
extensors in the limb contralateral to the trained underlies some forms of motor learning (Rioult-Pedotti
muscles when there was a 30% increase in MVC in et al. 1998, 2000), and inferences about the induction

 2011 The Authors


Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x 133
Neural adaptations to strength training Æ T J Carroll et al. Acta Physiol 2011, 202, 119–140

of LTP can be drawn in humans through rTMS rest or during weak contractions. This suggests that if
protocols that mimic LTP and long-term depression training changes the properties of cortical and spinal
(LTD). The principle that allows rTMS approaches to circuits projecting to agonist muscles, then the changes
provide inferences about the induction of LTP- or LTD- may be too subtle to be detected with MEP approaches
like changes in synaptic efficacy is that of ‘metaplastic- in many contexts. Despite this, there is some evidence
ity’, or ‘plasticity of plasticity’ (see Abraham & Bear from TMS studies that the responsiveness of motoneu-
1996). This principle holds that the effects of any given rons during high force contractions is reduced after
plasticity inducing protocol will be regulated depending strength training, that disruption of normal corticosp-
on the prior history of synaptic activity. The functional inal processing after training reduces strength gains, and
effect of this process is to maintain synaptic efficacy that training increases the capacity of the motor cortex
within an appropriate range. Thus, the effect of a to drive the motoneurons during maximal effort. In
protocol that typically enhances corticospinal respon- order to advance our understanding of the neural
siveness will be reduced if it is applied when the targeted adaptations to strength training, studies that target
circuits have experienced prior LTP (Ziemann et al. specific sites of modulation via converging techniques
2004, 2006). Accordingly, the effects of a protocol that (e.g. TMS vs. cervico-medullary stimulation) or condi-
typically suppresses corticospinal excitability will be tioning-test approaches are required. Close attention to
enhanced by prior induction of LTP. A series of studies the functional context under which the methods are
have used this general approach to argue that a motor applied will also be essential. Finally, rTMS protocols
learning training protocol that is not dissimilar to have potential to provide information regarding the
strength training (i.e. multiple ballistic contractions that nature of any synaptic plasticity that might be induced
increase peak joint acceleration) causes LTP-like adap- by strength training.
tations in the motor cortex (Ziemann et al. 2004, Stefan
et al. 2006, Rosenkranz et al. 2007). These studies used
Imaging and EEG studies
a technique known as paired associative stimulation
(PAS), which involves the application of peripheral Functional magnetic resonance imaging is a non-inva-
nerve stimulation (to activate muscle afferents) such sive technique that detects regional changes in cortical
that the afferent volley arrives near simultaneously at blood flow in response to a task measured as the blood
the motor cortex with TMS (Stefan et al. 2000). oxygenation level dependent (BOLD) contrast or ratio
Approximately 100 stimuli pairs are needed to modu- of oxygenated to deoxygenate haemoglobin (see Logo-
late the synaptic efficacy of cortical synapses for an thetis 2003 for review). Changes in BOLD response are
extended period (see Ziemann et al. 2008 for review). If interpreted as changes in ‘neural activation’ in fMRI
the afferent volley arrives at the cortex prior to TMS studies but are not direct measures of synaptic activity
[i.e. inter-stimulus interval (ISI) of 10 ms for hand or neuronal action potentials. While the spatial resolu-
muscles; PAS10], long lasting depression occurs. If the tion of fMRI is in the order of a few millimetres,
afferent volley arrives coincident with, or just subse- because the signal is sensitive to vascular changes and
quent to the TMS (i.e. ISI of 25 ms for hand muscles; not neuronal activity, the temporal resolution is in the
PAS25), long lasting facilitation results. PAS protocols order of a few seconds.
induce effects that share many characteristics with LTP/ Using fMRI, Farthing et al. (2007) examined changes
LTD; the changes in excitability last for a similar in brain activation following 6 weeks of unilateral
duration (30–60 min), they are dependent on normal strength training consisting of a maximum isometric
N-methyl-D-aspartate receptor function, they are spe- ulnar deviation task. The authors reported widespread
cific to the circuits targeted by stimulation, and are changes in the activation profile in both the contralat-
reversible by protocols that have opposite effects on eral (trained) and ipsilateral (untrained) cortices in
synaptic efficacy (see Ziemann et al. 2008 for review). parallel with increased strength in both the trained
Although there are many technical challenges in apply- (right) and untrained (left) limb. There are, however,
ing PAS and other rTMS techniques that can induce several limitations to the interpretation of changes in
synaptic plasticity to strength training studies, the brain activation from fMRI (see Poldrack 2000, Kelly
potential to establish whether strength training induces & Garavan 2005 for review). Firstly, it is not possible to
LTP-like changes in the motor cortex dictate that the determine the extent to which areas of activation
experiments should be attempted. detected by fMRI are functionally related to and share
a causal relationship with behaviour, or merely repre-
sent correlated activity associated with task-irrelevant
Moving forward with TMS studies
processing. For example, changes in behavioural output
There is a lack of consistency in the results of studies or cognitive awareness of the imaging process can
that assessed the effect of strength training on MEPs at influence brain activation measured with fMRI. Infor-

 2011 The Authors


134 Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x
Acta Physiol 2011, 202, 119–140 T J Carroll et al. Æ Neural adaptations to strength training
mation regarding the mechanisms of cortical changes resolution (i.e. over fMRI, for example). However, the
associated with strength training might be best achieved spatial resolution of EEG is much lower than most
by testing specific predictions regarding changes in brain imaging techniques, such as fMRI. Surface negative
activity, based on the known functional roles of cortical potentials detected at the scalp around the time of
areas and likely time course of the hypothesized effects, voluntary movement are referred to as movement-
rather than simply documenting changes in cortical related cortical potentials (MRCP). MRCPs reflect the
activity in response to training. For example, in assess- summed excitatory post-synaptic potentials of apical
ing transfer of motor skill from one limb to the other, dendrites and are related to the preparation and
Perez et al. (2007) found that the level of pre-training execution of self-initiated movement (see Shibasaki &
fMRI activity specific to the ventrolateral posterior Hallett 2006 for review). The MRCP waveform has an
thalamic (VLp) nucleus predicted successful future onset 1–2 s prior to movement onset and can be divided
intermanual transfer. These observations are consistent into three consecutive phases referred to as (1) the
with what is known about the associated connections Bereitschaftspotential (preparation), (2) motor execu-
between VLp, cerebellum and supplementary motor tion and (3) movement-monitoring potentials. The
area, which are all implicated in error prediction and amplitude of each component is a function of the
correction (Butler et al. 2000, Sakai et al. 2002, Died- number of active neurons, their synchrony and rate of
richsen et al. 2005). Secondly, using fMRI, it is not discharge (Siemionow et al. 2000). The later MRCP
possible to dissociate activity that is excitatory in nature components have been correlated with force, RFD, and
from that which is inhibitory. Therefore, the overall associated EMG amplitude for both elbow flexion
activation profiles observed in fMRI studies are a (Siemionow et al. 2000), and plantar flexion move-
combination of excitatory and inhibitory neural pro- ments (do Nascimento et al. 2005) suggesting that
cesses and their associated BOLD signal. As such, it is MRCP is indicative of the level of muscle activation.
often unclear how to interpret increases in ‘activity’ A 3-week resistance training program, with increases
based on fMRI. in MVC, RFD and EMG in the trained leg extensor
Another important consideration in interpreting stud- muscles, resulted in changes in MRCP amplitude at
ies that examine the cortical changes associated with several motor electrode sites (Falvo et al. 2010).
strength training is the time course of the changes. For For repetitive submaximal leg extensions, Falvo et al.
example, observed changes in M1 representation with (2010) observed an attenuation of MRCP amplitude.
motor learning (Pascual-Leone et al. 1994, Karni et al. These findings support the hypothesis that by increasing
1995) are dependent on the time course of measure- muscle strength, comparable motor tasks can be
ment. Xiong et al. (2009) found increases in M1 activity performed with a lower level of neural activation (e.g.
after 2 weeks of motor practice, but activity levels Carroll et al. 2001, Carson 2006). As the force gener-
returned to baseline by week four. It is likely that, if ating capacity of individual motor units increases with
there are cortical changes associated with strength training, fewer motor neurons are required in order to
training, they are also time dependent. As mentioned accomplish a given submaximal task. Such a reduction
in an earlier section, alterations in cortical activation in motor unit recruitment might be reflected in dimin-
which might contribute to early increases in strength ished cortical activation.
might be associated with generating an increase in Finally, frequency-domain coupling between cortical
agonist and synergist activation while inhibiting antag- activity and spinal motoneuron firing can be captured in
onist activation. Alternatively, strength training that is the EEG–EMG coherence signal (see Conway et al.
sufficient to result in enhancements in peripheral force 1995). It has recently been shown that corticomuscular
generating capacity, might be expected to result in less coherence is lower for strength trained than for non-
cortical activation to produce a given level of force (i.e. trained individuals (Ushiyama et al. 2010), which sug-
because fewer motor units would be require activation). gests that the method may have potential for studying
It remains uncertain whether fMRI has sufficient the corticospinal responses to strength training.
resolution to detect potentially subtle changes in brain
organization underlying enhanced muscular coordina-
General summary and conclusions
tion in the presence of peripheral changes.
Another technique for monitoring brain activity is In this review, we summarized the available research in
EEG, a non-invasive technique that records electrical which neurophysiological methods were applied to
activity in the brain. Surface EEG electrodes measure study neural adaptations to strength training. Despite
voltage changes on the scalp that reflect ion flow caused a considerable body of work spanning over 30 years,
by excitatory and inhibitory post-synaptic potentials. As there remain many inconsistencies in the data and many
EEG is derived from post-synaptic potentials in cortical gaps in our understanding. One possible interpretation
neurons, it has the advantage of excellent temporal of the inconsistent results is that variations in the

 2011 The Authors


Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x 135
Neural adaptations to strength training Æ T J Carroll et al. Acta Physiol 2011, 202, 119–140

training task and muscle groups targeted fundamentally Bigland-Ritchie, B., Jones, D.A. & Woods, J.J. 1979. Excita-
alter the nature of the neural adaptations produced. tion frequency and muscle fatigue: electrical responses dur-
However, we do not favour this interpretation. In our ing human voluntary and stimulated contractions. Exp
opinion, although there is likely to be a continuum of Neurol 64, 414–427.
Bostock, H. & Grafe, P. 1985. Activity-dependent excitability
changes that will depend upon the precise details of the
changes in normal and demyelinated rat spinal root axons.
training and the characteristics of the involved neuro-
J Physiol 365, 239–257.
muscular components, we expect that it should ulti- Brocke, J., Irlbacher, K., Hauptmann, B., Voss, M. & Brandt,
mately be possible to identify some general principles of S.A. 2005. Transcranial magnetic and electrical stimulation
neural adaptation that are applicable to all forms of compared: does TES activate intracortical neuronal circuits?
strength training that involve systematic repetition of Clin Neurophysiol 116, 2748–2756.
actions requiring strong neural drive. Clearly, if these Burke, D. & Gandevia, S.C. 1999. Properties of human
general principles exist, they cannot be identified on the peripheral nerves: implications for studies of human motor
basis of current data. We conclude that in order to control. In: M.D. Binder (ed.) Peripheral and Spinal Mech-
advance knowledge in this field, studies should be anisms in the Neural Control of Movement, pp. 427–435.
designed to rigorously account for the limitations of the Elsevier Science, Amsterdam.
Burke, D., Gandevia, S.C. & McKeon, B. 1984. Mono-syn-
available techniques, and be specifically targeted to
aptic and oligosynaptic contributions to human ankle jerk
address important conceptual questions. Careful con-
and H-reflex. J Neurophysiol 52, 435–448.
sideration of the likely time course of mechanisms to be Butler, E.G., Bourke, D.W. & Horne, M.K. 2000. The activity
studied, and the functional context in which they are of primate ventrolateral thalamic neurones during motor
investigated will be critical. adaptation. Exp Brain Res 133, 514–531.
Capaday, C. 1997. Neurophysiological methods for studies of
the motor system in freely moving human subjects. J Neu-
Conflict of interest
rosci Methods 74, 201–218.
None. Carroll, T.J., Barry, B., Riek, S. & Carson, R.G. 2001. Resis-
tance training enhances the stability of sensorimotor coor-
This work was supported by the Australian Research Council dination. Proc Biol Sci 268, 221–227.
(TJC, SR) and the National Health and Medical Research Carroll, T.J., Riek, S. & Carson, R.G. 2002. The sites of neural
Council (JGS). The authors thank Dr Michael Lee and adaptation induced by resistance training in humans.
Professors Simon Gandevia and Richard Carson for discussions J Physiol 544, 641–652.
that helped shape many of the ideas in this paper. Carroll, T.J., Barton, J., Hsu, M. & Lee, M. 2009. The effect of
strength training on the force of twitches evoked by
corticospinal stimulation in humans. Acta Physiol 197, 161–
References
173.
Aagaard, P., Simonsen, E.B., Andersen, J.L., Magnusson, P. & Carson, R.G. 2006. Changes in muscle coordination with
Dyhre-Poulsen, P. 2002. Neural adaptation to resistance training. J Appl Physiol 101, 1506–1513.
training: changes in evoked V-wave and H-reflex responses. Chen, R., Classen, J., Gerloff, C., Celnik, P., Wassermann,
J Appl Physiol 92, 2309–2318. E.M., Hallett, M. & Cohen, L.G. 1997. Depression of motor
Abraham, W.C. & Bear, M.F. 1996. Metaplasticity: the plas- cortex excitability by low-frequency transcranial magnetic
ticity of synaptic plasticity. Trends Neurosci 19, 126–130. stimulation. Neurology 48, 1398–1403.
Adkins, D.L., Boychuk, J., Remple, M.S. & Kleim, J.A. 2006. Christie, A. & Kamen, G. 2006. Doublet discharges in moto-
Motor training induces experience-specific patterns of plas- neurons of young and older adults. J Neurophysiol 95,
ticity across motor cortex and spinal cord. J Appl Physiol 2787–2795.
101, 1776–1782. Christie, A. & Kamen, G. 2010. Short-term training adapta-
Adrian, E.D. & Bronk, D.W. 1929. The discharge of impulses tions in maximal motor unit firing rates and afterhyperpo-
in motor nerve fibres: part II. The frequency of discharge in larization duration. Muscle Nerve 41, 651–660.
reflex and voluntary contractions. J Physiol 67, i3–i151. Conway, B.A., Halliday, D.M., Farmer, S.F., Shahani, U.,
Barker, A.T. 1991. An introduction to the basic principles of Maas, P., Weir, A.I. & Rosenberg, J.R. 1995. Synchroniza-
magnetic nerve-stimulation. J Clin Neurophysiol 8, 26–37. tion between motor cortex and spinal motoneuronal pool
Barker, A.T., Jalinous, R. & Freeston, I.L. 1985. Non-invasive during the performance of a maintained motor task in man.
magnetic stimulation of human motor cortex. Lancet 1, J Physiol 489, 917–924.
1106–1107. Dartnall, T.J., Nordstrom, M.A. & Semmler, J.G. 2008. Motor
Bawa, P. & Calancie, B. 1983. Repetitive doublets in human unit synchronization is increased in biceps brachii after
flexor carpi radialis muscle. J Physiol 339, 123–132. exercise-induced damage to elbow flexor muscles. J Neuro-
Beck, S., Taube, W., Gruber, M., Amtage, F., Gollhofer, A. & physiol 99, 1008–1019.
Schubert, M. 2007. Task-specific changes in motor evoked Day, B.L., Dressler, D., Maertens de Noordhout, A., Marsden,
potentials of lower limb muscles after different training C.D., Nakashima, K., Rothwell, J.C. & Thompson, P.D.
interventions. Brain Res 1179, 51–60. 1989. Electric and magnetic stimulation of human motor

 2011 The Authors


136 Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x
Acta Physiol 2011, 202, 119–140 T J Carroll et al. Æ Neural adaptations to strength training
cortex: surface EMG and single motor unit responses. synaptic inputs to motoneurones studied during voluntary
J Physiol 412, 449–473. isometric contraction in man. J Physiol 470, 127–155.
De Luca, C.J., LeFever, R.S., McCue, M.P. & Xenakis, A.P. Farmer, S.F., Halliday, D.M., Conway, B.A., Stephens, J.A. &
1982. Control scheme governing concurrently active human Rosenberg, J.R. 1997. A review of recent applications of
motor units during voluntary contractions. J Physiol 329, cross-correlation methodologies to human motor unit
129–142. recording. J Neurosci Methods 74, 175–187.
De Luca, C.J., Adam, A., Wotiz, R., Gilmore, L.D. & Nawab, Farthing, J.P., Borowsky, R., Chilibeck, P.D., Binsted, G. &
S.H. 2006. Decomposition of surface EMG signals. J Neu- Sarty, G.E. 2007. Neuro-physiological adaptations associ-
rophysiol 96, 1646–1657. ated with cross-education of strength. Brain Topogr 20,
Del Balso, C. & Cafarelli, E. 2007. Adaptations in the acti- 77–88.
vation of human skeletal muscle induced by short-term iso- Fimland, M.S., Helgerud, J., Gruber, M., Leivseth, G. & Hoff,
metric resistance training. J Appl Physiol 103, 402–411. J. 2009a. Functional maximal strength training induces
Del Valle, A. & Thomas, C.K. 2005. Firing rates of motor units neural transfer to single-joint tasks. Eur J Appl Physiol 107,
during strong dynamic contractions. Muscle Nerve 32, 316– 21–29.
325. Fimland, M.S., Helgerud, J., Solstad, G.M., Iversen, V.M.,
Di Lazzaro, V., Oliviero, A., Profice, P., Saturno, E., Pilato, F., Leivseth, G. & Hoff, J. 2009b. Neural adaptations under-
Insola, A., Mazzone, P., Tonali, P. & Rothwell, J.C. 1998a. lying cross-education after unilateral strength training. Eur J
Comparison of descending volleys evoked by transcranial Appl Physiol 107, 723–730.
magnetic and electric stimulation in conscious humans. Fitzgerald, P.B., Fountain, S. & Daskalakis, Z.J. 2006. A
Electroencephalogr Clin Neurophysiol 109, 397–401. comprehensive review of the effects of rTMS on motor
Di Lazzaro, V., Restuccia, D., Oliviero, A., Profice, P., Ferrara, cortical excitability and inhibition. Clin Neurophysiol 117,
L., Insola, A., Mazzone, P., Tonali, P. & Rothwell, J.C. 2584–2596.
1998b. Effects of voluntary contraction on descending vol- Fling, B.W., Christie, A. & Kamen, G. 2009. Motor unit syn-
leys evoked by transcranial stimulation in conscious humans. chronization in FDI and biceps brachii muscles of strength-
J Physiol 508, 625–633. trained males. J Electromyogr Kinesiol 19, 800–809.
Di Lazzaro, V., Oliviero, A., Pilato, F., Saturno, E., Dileone, Frigon, A., Carroll, T.J., Jones, K.E., Zehr, E.P. & Collins,
M., Mazzone, P., Insola, A., Tonali, P.A. & Rothwell, J.C. D.F. 2007. Ankle position and voluntary contraction alter
2004. The physiological basis of transcranial motor cortex maximal M waves in soleus and tibialis anterior. Muscle
stimulation in conscious humans. Clin Neurophysiol 115, Nerve 35, 756–766.
255–266. Gandevia, S.C. 2001. Spinal and supraspinal factors in human
Diedrichsen, J., Hashambhoy, Y., Rane, T. & Shadmehr, R. muscle fatigue. Physiol Rev 81, 1725–1789.
2005. Neural correlates of reach errors. J Neurosci 25, Gandevia, S.C., Allen, G.M., Butler, J.E. & Taylor, J.L. 1996.
9919–9931. Supraspinal factors in human muscle fatigue: evidence for
Doherty, T.J. & Brown, W.F. 1994. A method for the longi- suboptimal output from the motor cortex. J Physiol 490,
tudinal study of human thenar motor units. Muscle Nerve 529–536.
17, 1029–1036. Gandevia, S.C., Petersen, N., Butler, J.E. & Taylor, J.L. 1999.
Duclay, J., Martin, A., Robbe, A. & Pousson, M. 2008. Spinal Impaired response of human motoneurones to corticospinal
reflex plasticity during maximal dynamic contractions after stimulation after voluntary exercise. J Physiol 521, 749–759.
eccentric training. Med Sci Sports Exerc 40, 722–734. Garland, S.J. & Griffin, L. 1999. Motor unit double dis-
Ekblom, M.M.N. 2010. Improvements in dynamic plantar charges: statistical anomaly or functional entity? Can J Appl
flexor strength after resistance training are associated with Physiol 24, 113–130.
increased voluntary activation and V-to-M ratio. J Appl Geertsen, S.S., Lundbye-Jensen, J. & Nielsen, J.B. 2008. In-
Physiol 109, 19–26. creased central facilitation of antagonist reciprocal inhibition
Enoka, R.M. 1996. Eccentric contractions require unique at the onset of dorsiflexion following explosive strength
activation strategies by the nervous system. J Appl Physiol training. J Appl Physiol 105, 915–922.
81, 2339–2346. Gondin, J., Duclay, J. & Martin, A. 2006. Soleus- and gas-
Enoka, R.M., Robinson, G.A. & Kossev, A.R. 1988. A stable, trocnemii-evoked V-wave responses increase after neuro-
selective electrode for recording single motor-unit potentials muscular electrical stimulation training. J Neurophysiol 95,
in humans. Exp Neurol 99, 761–764. 3328–3335.
Falvo, M.J., Sirevaag, E.J., Rohrbaugh, J.W. & Earhart, G.M. Gooch, C.L. & Harati, Y. 1997. Longitudinal tracking of the
2010. Resistance training induces supraspinal adaptations: same single motor unit in amyotrophic lateral sclerosis.
evidence from movement-related cortical potentials. Eur J Muscle Nerve 20, 511–513.
Appl Physiol 109, 923–933. Griffin, L. & Cafarelli, E. 2007. Transcranial magnetic stim-
Farina, D., Holobar, A., Merletti, R. & Enoka, R.M. 2010. ulation during resistance training of the tibialis anterior
Decoding the neural drive to muscles from the surface elec- muscle. J Electromyogr Kinesiol 17, 446–452.
tromyogram. Clin Neurophysiol 121, 1616–1623. Griffin, L., Painter, P.E., Wadhwa, A. & Spirduso, W.W. 2009.
Farmer, S.F., Bremner, F.D., Halliday, D.M., Rosenberg, J.R. Motor unit firing variability and synchronization during
& Stephens, J.A. 1993. The frequency content of common short-term light-load training in older adults. Exp Brain Res
197, 337–345.

 2011 The Authors


Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x 137
Neural adaptations to strength training Æ T J Carroll et al. Acta Physiol 2011, 202, 119–140

Gydikov, A., Kossev, A., Trayanova, N. & Radicheva, N. Lagerquist, O., Zehr, E.P. & Docherty, D. 2006. Increased
1986. Selective recording of motor unit potentials. Electro- spinal reflex excitability is not associated with neural plas-
myogr Clin Neurophysiol 26, 273–281. ticity underlying the cross-education effect. J Appl Physiol
Hashimoto, R. & Rothwell, J.C. 1999. Dynamic changes in 100, 83–90.
corticospinal excitability during motor imagery. Exp Brain Laidlaw, D.H., Kornatz, K.W., Keen, D.A., Suzuki, S. &
Res 125, 75–81. Enoka, R.M. 1999. Strength training improves the steadiness
Holtermann, A., Roeleveld, K., Engstrom, M. & Sand, T. of slow lengthening contractions performed by old adults.
2007. Enhanced H-reflex with resistance training is related J Appl Physiol 87, 1786–1795.
to increased rate of force development. Eur J Appl Physiol Lee, M. & Carroll, T.J. 2005. The amplitude of Mmax in human
101, 301–312. wrist flexors varies during different muscle contractions
Hortobagyi, T., Richardson, S.P., Lomarev, M., Shamim, E., despite constant posture. J Neurosci Methods 149, 95–100.
Meunier, S., Russman, H., Dang, N. & Hallett, M. 2009. Lee, M., Gandevia, S.C. & Carroll, T.J. 2008. Cortical vol-
Chronic low-frequency rTMS of primary motor cortex untary activation can be reliably measured in human wrist
diminishes exercise training-induced gains in maximal vol- extensors using transcranial magnetic stimulation. Clin
untary force in humans. J Appl Physiol 106, 403–411. Neurophysiol 119, 1130–1138.
Hultborn, H. & Nielsen, J.B. 1998. Modulation of transmitter Lee, M., Gandevia, S.C. & Carroll, T.J. 2009. Short-term
release from Ia afferents by their preceding activity – a strength training does not change cortical voluntary activa-
‘postactivation depression’. In: P. Rudomin & L. Mendell tion. Med Sci Sports Exerc 41, 1452–1460.
(eds) Presynaptic Inhibition and Neural Control, pp. 178– Lee, M., Hinder, M.R., Gandevia, S.C. & Carroll, T.J. 2010.
191. Oxford University Press, New York. The ipsilateral motor cortex contributes to cross-limb
Jensen, J.L., Marstrand, P.C.D. & Nielsen, J.B. 2005. Motor transfer of performance gains after ballistic motor practice.
skill training and strength training are associated with dif- J Physiol 588, 201–212.
ferent plastic changes in the central nervous system. J Appl Logothetis, N.K. 2003. MR imaging in the non-human pri-
Physiol 99, 1558–1568. mate: studies of function and of dynamic connectivity. Curr
Kamen, G. & Knight, C.A. 2004. Training-related adaptations Opin Neurobiol 13, 630–642.
in motor unit discharge rate in young and older adults. Macefield, V.G., Gandevia, S.C., Bigland-Ritchie, B., Gorman,
J Gerontol A Biol Sci Med Sci 59, 1334–1338. R.B. & Burke, D. 1993. The firing rates of human moto-
Kamen, G., Sison, S.V., Du, C.C. & Patten, C. 1995. Motor neurones voluntarily activated in the absence of muscle
unit discharge behavior in older adults during maximal- afferent feedback. J Physiol 471, 429–443.
effort contractions. J Appl Physiol 79, 1908–1913. Macefield, V.G., Fuglevand, A.J. & BiglandRitchie, B. 1996.
Karni, A., Meyer, G., Jezzard, P., Adams, M.M., Turner, R. & Contractile properties of single motor units in human toe
Ungerleider, L.G. 1995. Functional MRI evidence for adults extensors assessed by intraneural motor axon stimulation.
motor cortex plasticity during motor skiill learning. Nature J Neurophysiol 75, 2509–2519.
377, 155–158. Maffiuletti, N.A., Pensini, M., Scaglioni, G., Ferri, A., Ballay,
Keen, D.A., Yue, G.H. & Enoka, R.M. 1994. Training-related Y. & Martin, A. 2003. Effect of electromyostimulation
enhancement in the control of motor output in elderly training on soleus and gastrocnemii H- and T-reflex prop-
humans. J Appl Physiol 77, 2648–2658. erties. Eur J Appl Physiol 90, 601–607.
Kelly, A.M. & Garavan, H. 2005. Human functional neuroi- Magladery, J.W. & McDougal, D.B. 1950. Electrophysiologi-
maging of brain changes associated with practice. Cereb cal studies of nerve and reflex activity in normal man.1.
Cortex 15, 1089–1102. Identification of certain reflexes in the electromyogram and
Kidgell, D.J. & Pearce, A.J. 2010. Corticospinal properties the conduction velocity of peripheral nerve fibres. Bull Johns
following short-term strength training of an intrinsic hand Hopkins Hosp 86, 265–290.
muscle. Hum Mov Sci 29, 631–641. Magladery, J.W., Porter, W.E., Park, A.M. & Teasdall, R.D.
Kidgell, D.J., Sale, M.V. & Semmler, J.G. 2006. Motor unit 1951. Electrophysiological studies of nerve and reflex
synchronization measured by cross-correlation is not influ- activity in normal man. IV. The two-neurone reflex and
enced by short-term strength training of a hand muscle. Exp identification of certain action potentials from
Brain Res 175, 745–753. spinal roots and cord. Bull Johns Hopkins Hosp 88, 499–
Klass, M., Baudry, S. & Duchateau, J. 2008. Age-related de- 519.
cline in rate of torque development is accompanied by lower Marchand-Pauvert, V., Nicolas, G., Burke, D. & Pierrot-Des-
maximal motor unit discharge frequency during fast con- eilligny, E. 2002. Suppression of the H reflex in humans by
tractions. J Appl Physiol 104, 739–746. disynaptic autogenetic inhibitory pathways activated by the
Knight, C.A. & Kamen, G. 2004. Enhanced motor unit rate test volley. J Physiol 542, 963–976.
coding with improvements in a force-matching task. J Elec- Matthews, P.B. 1999. The effect of firing on the excitability of
tromyogr Kinesiol 14, 619–629. a model motoneurone and its implications for cortical
Kornatz, K.W., Christou, E.A. & Enoka, R.M. 2005. Practice stimulation. J Physiol 518, 867–882.
reduces motor unit discharge variability in a hand muscle McHugh, M.P. 2003. Recent advances in the understanding of
and improves manual dexterity in old adults. J Appl Physiol the repeated bout effect: the protective effect against muscle
98, 2072–2080. damage from a single bout of eccentric exercise. Scand J Med
Sci Sports 13, 88–97.

 2011 The Authors


138 Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x
Acta Physiol 2011, 202, 119–140 T J Carroll et al. Æ Neural adaptations to strength training
Merletti, R., Farina, D. & Gazzoni, M. 2003. The linear Perez, M.A., Tanaka, S., Wise, S.P., Sadato, N., Tanabe, H.C.,
electrode array: a useful tool with many applications. Willingham, D.T. & Cohen, L.G. 2007. Neural substrates of
J Electromyogr Kinesiol 13, 37–47. intermanual transfer of a newly acquired motor skill. Curr
Merletti, R., Holobar, A. & Farina, D. 2008. Analysis of motor Biol 17, 1896–1902.
units with high-density surface electromyography. J Elec- Petersen, N.T., Pyndt, H.S. & Nielsen, J.B. 2003. Investigating
tromyogr Kinesiol 18, 879–890. human motor control by transcranial magnetic stimulation.
Merton, P.A. 1954. Voluntary strength and fatigue. J Physiol Exp Brain Res 152, 1–16.
123, 553–564. Pierrot-Deseilligny, E. & Burke, D. 2005. The Circuitry of the
Miller, K.J., Garland, S.J., Ivanova, T. & Ohtsuki, T. 1996. Human Spinal Cord: Its Role in Motor Control and Move-
Motor-unit behavior in humans during fatiguing arm ment Disorders. Cambridge University Press, Cambridge.
movements. J Neurophysiol 75, 1629–1636. Poldrack, R.A. 2000. Imaging brain plasticity: conceptual and
Milner-Brown, H.S., Stein, R.B. & Yemm, R. 1973. The con- methodological issues – a theoretical review. Neuroimage
tractile properties of human motor units during voluntary 12, 1–13.
isometric contractions. J Physiol 228, 285–306. Porter, R. & Lemon, R.N. 1993. Corticospinal Function and
Milner-Brown, H.S., Stein, R.B. & Lee, R.G. 1975. Synchro- Voluntary Movement. Oxford University Press, Oxford.
nization of human motor units-possible roles of exercise and Pucci, A.R., Griffin, L. & Cafarelli, E. 2006. Maximal motor
supraspinal reflexes. Electroencephalogr Clin Neurophysiol unit firing rates during isometric resistance training in men.
38, 245–254. Exp Physiol 91, 171–178.
Misiaszek, J.E. 2003. The H-reflex as a tool in neurophysiol- Reis, J., Swayne, O.B., Vandermeeren, Y., Camus, M., Dim-
ogy: its limitations and uses in understanding nervous system yan, M.A., Harris-Love, M., Perez, M.A., Ragert, P., Roth-
function. Muscle Nerve 28, 144–160. well, J.C. & Cohen, L.G. 2008. Contribution of transcranial
Mogyoros, I., Kiernan, M.C. & Burke, D. 1996. Strength- magnetic stimulation to the understanding of cortical
duration properties of human peripheral nerve. Brain 119, mechanisms involved in motor control. J Physiol 586, 325–
439–447. 351.
Moritz, C.T., Barry, B.K., Pascoe, M.A. & Enoka, R.M. 2005. Rioult-Pedotti, M.S., Friedman, D., Hess, G. & Donoghue, J.P.
Discharge rate variability influences the variation in force 1998. Strengthening of horizontal cortical connections fol-
fluctuations across the working range of a hand muscle. lowing skill learning. Nat Neurosci 1, 230–234.
J Neurophysiol 93, 2449–2459. Rioult-Pedotti, M.S., Friedman, D. & Donoghue, J.P. 2000.
Mottram, C.J., Jakobi, J.M., Semmler, J.G. & Enoka, R.M. Learning-induced LTP in neocortex. Science 290, 533–536.
2005. Motor-unit activity differs with load type during a Rosenkranz, K., Kacar, A. & Rothwell, J.C. 2007. Differential
fatiguing contraction. J Neurophysiol 93, 1381–1392. modulation of motor cortical plasticity and excitability in
do Nascimento, O.F., Nielsen, K.D. & Voigt, M. 2005. Rela- early and late phases of human motor learning. J Neurosci
tionship between plantar-flexor torque generation and the 27, 12058–12066.
magnitude of the movement-related potentials. Exp Brain Ruohonen, J. & Ilmoniemi, R.J. 1999. Modeling of the stim-
Res 160, 154–165. ulating field generation in TMS. In: W. Paulus, M. Hallett,
Nawab, S.H., Chang, S.S. & De Luca, C.J. 2010. High-yield P.M. Rossini & J.C. Rothwell (eds) Transcranial Magnetic
decomposition of surface EMG signals. Clin Neurophysiol Stimulation. Electroencephalography and Clinical Neuro-
121, 1602–1615. physiology (Supplement 51), pp. 30–40. Elsevier Science,
Nicol, C., Avela, J. & Komi, P.V. 2006. The stretch-shortening Amsterdam.
cycle : a model to study naturally occurring neuromuscular Sakai, S.T., Inase, M. & Tanji, J. 2002. The relationship
fatigue. Sports Med 36, 977–999. between MI and SMA afferents and cerebellar and pallidal
Nielsen, J., Petersen, N., Deuschl, G. & Ballegaard, M. 1993. efferents in the macaque monkey. Somatosens Mot Res 19,
Task-related changes in the effect of magnetic brain 139–148.
stimulation on spinal neurones in man. J Physiol 471, 223– Sale, D.G., McComas, A.J., Macdougall, J.D. & Upton,
243. A.R.M. 1982. Neuromuscular adaptation in human thenar
Oya, T., Riek, S. & Cresswell, A.G. 2009. Recruitment and muscles following strength training and immobilization.
rate coding organisation for soleus motor units across entire J Appl Physiol 53, 419–424.
range of voluntary isometric plantar flexions. J Physiol 587, Sale, D.G., Macdougall, J.D., Upton, A.R.M. & McComas,
4737–4748. A.J. 1983. Effects of strength training upon motoneuron
Pascual-Leone, A., Grafman, J. & Hallett, M. 1994. Modula- excitability in man. Med Sci Sports Exerc 15, 57–62.
tion of cortical motor output maps during development of Scaglioni, G., Ferri, A., Minetti, A.E., Martin, A., Van Hoecke,
implicit and explicit knowledge. Science 263, 1287–1289. J., Capodaglio, P., Sartorio, A. & Narici, M.V. 2002. Plantar
Pasquet, B., Carpentier, A. & Duchateau, J. 2006. Specific flexor activation capacity and H reflex in older adults:
modulation of motor unit discharge for a similar change in adaptations to strength training. J Appl Physiol 92, 2292–
fascicle length during shortening and lengthening contrac- 2302.
tions in humans. J Physiol 577, 753–765. Schneider, J., Rau, G. & Silny, J. 1989. A noninvasive EMG
Patten, C., Kamen, G. & Rowland, D.M. 2001. Adaptations in technique for investigating the excitation propagation in
maximal motor unit discharge rate to strength training in single motor units. Electromyogr Clin Neurophysiol 29,
young and older adults. Muscle Nerve 24, 542–550. 273–280.

 2011 The Authors


Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x 139
Neural adaptations to strength training Æ T J Carroll et al. Acta Physiol 2011, 202, 119–140

Schubert, M., Beck, S., Taube, W., Amtage, F., Faist, M. & flexors with motor cortical stimulation. J Appl Physiol 97,
Gruber, M. 2008. Balance training and ballistic strength 236–242.
training are associated with task-specific corticospinal Tracy, B.L., Maluf, K.S., Stephenson, J.L., Hunter, S.K. &
adaptations. Eur J Neurosci 27, 2007–2018. Enoka, R.M. 2005. Variability of motor unit discharge and
Semmler, J.G. & Nordstrom, M.A. 1998. Motor unit discharge force fluctuations across a range of muscle forces in older
and force tremor in skill- and strength-trained individuals. adults. Muscle Nerve 32, 533–540.
Exp Brain Res 119, 27–38. Ugawa, Y., Rothwell, J.C., Day, B.L., Thompson, P.D. &
Semmler, J.G., Kornatz, K.W., Dinenno, D.V., Zhou, S. & Marsden, C.D. 1991. Precutaneous electric-stimulation of
Enoka, R.M. 2002. Motor unit synchronisation is enhanced corticospinal pathways at the level of the pyramidal decus-
during slow lengthening contractions of a hand muscle. sation in humans. Ann Neurol 29, 418–427.
J Physiol 545, 681–695. Upton, A.R., McComas, A.J. & Sica, R.E. 1971. Potentiation
Semmler, J.G., Sale, M.V., Meyer, F.G. & Nordstrom, M.A. of ‘late’ responses evoked in muscles during effort. J Neurol
2004. Motor-unit coherence and its relation with Neurosurg Psychiatry 34, 699–711.
synchrony are influenced by training. J Neurophysiol 92, Ushiyama, J., Takahashi, Y. & Ushiba, J. 2010. Muscle
3320–3331. dependency of corticomuscular coherence in upper and
Shibasaki, H. & Hallett, M. 2006. What is the Bereitschafts- lower limb muscles and training-related alterations in
potential? Clin Neurophysiol 117, 2341–2356. ballet dancers and weightlifters. J Appl Physiol 109, 1086–
Sidhu, S.K., Bentley, D.J. & Carroll, T.J. 2009. Cortical 1095.
voluntary activation of the human knee extensors can be Vagg, R., Mogyoros, I., Kiernan, M.C. & Burke, D. 1998.
reliably estimated using transcranial magnetic stimulation. Activity-dependent hyperpolarization of human motor axons
Muscle Nerve 39, 186–196. produced by natural activity. J Physiol 507, 919–925.
Siemionow, V., Yue, G.H., Ranganathan, V.K., Liu, J.Z. & Van Cutsem, M., Duchateau, J. & Hainaut, K. 1998. Changes
Sahgal, V. 2000. Relationship between motor activity-re- in single motor unit behaviour contribute to the increase in
lated cortical potential and voluntary muscle activation. Exp contraction speed after dynamic training in humans.
Brain Res 133, 303–311. J Physiol 513, 295–305.
Stalberg, E., Nandedkar, S.D., Sanders, D.B. & Falck, B. 1996. Vila-Cha, C., Falla, D. & Farina, D. 2010. Motor unit
Quantitative motor unit potential analysis. J Clin Neuro- behavior during submaximal contractions following six
physiol 13, 401–422. weeks of either endurance or strength training. J Appl
Stefan, K., Kunesch, E., Cohen, L.G., Benecke, R. & Classen, J. Physiol 109, 1455–1466.
2000. Induction of plasticity in the human motor cortex by Xiong, J.H., Ma, L.S., Wang, B.Q., Narayana, S., Duff, E.P.,
paired associative stimulation. Brain 123, 572–584. Egan, G.F. & Fox, P.T. 2009. Long-term motor training
Stefan, K., Wycislo, M., Gentner, R., Schramm, A., Naumann, induced changes in regional cerebral blood flow in both task
M., Reiners, K. & Classen, J. 2006. Temporary occlusion of and resting states. Neuroimage 45, 75–82.
associative motor cortical plasticity by prior dynamic motor Yao, W.X., Fuglevand, A.J. & Enoka, R.M. 2000. Motor-unit
training. Cereb Cortex 16, 376–385. synchronization increases EMG amplitude and decreases
Stuart, M., Butler, J.E., Collins, D.F., Taylor, J.L. & Gandevia, force steadiness of simulated contractions. J Neurophysiol
S.C. 2002. The history of contraction of the wrist flexors can 83, 441–452.
change cortical excitability. J Physiol 545, 731–737. Yue, G., Fuglevand, A.J., Nordstrom, M.A. & Enoka, R.M.
Taylor, J.L. & Gandevia, S.C. 2004. Noninvasive stimulation of 1995. Limitations of the surface electromyography technique
the human corticospinal tract. J Appl Physiol 96, 1496–1503. for estimating motor unit synchronization. Biol Cybern 73,
Taylor, J.L., Petersen, N.T., Butler, J.E. & Gandevia, S.C. 223–233.
2002. Interaction of transcranial magnetic stimulation Zehr, E.P. 2002. Considerations for use of the Hoffmann reflex
and electrical transmastoid stimulation in human subjects. in exercise studies. Eur J Appl Physiol 86, 455–468.
J Physiol 541, 949–958. Zhou, P. & Rymer, W.Z. 2004. Factors governing the form of
Taylor, A.M., Christou, E.A. & Enoka, R.M. 2003. Multiple the relation between muscle force and the EMG: a simula-
features of motor-unit activity influence force fluctuations tion study. J Neurophysiol 92, 2878–2886.
during isometric contractions. J Neurophysiol 90, 1350– Ziemann, U., Iliac, T.V., Pauli, C., Meintzschel, F. & Ruge, D.
1361. 2004. Learning modifies subsequent induction of long-term
Thomas, C.K., Butler, J.E. & Zijdewind, I. 2002. Patterns of potentiation-like and long-term depression-like plasticity in
pathological firing in human motor units. In: S.C. Gandevia, human motor cortex. J Neurosci 24, 1666–1672.
U. Proske & D.G. Stuart (eds) Sensorimotor Control of Ziemann, U., Ilic, T.V. & Jung, P. 2006. Long-term potentia-
Movement and Posture, pp. 237–244. Kluwer Academic/ tion (LTP)-like plasticity and learning in human motor cor-
Plenum Publishers, New York. tex – investigations with transcranial magnetic stimulation
Todd, G., Taylor, J.L. & Gandevia, S.C. 2003. Measurement (TMS). Suppl Clin Neurophysiol 59, 19–25.
of voluntary activation of fresh and fatigued human muscles Ziemann, U., Paulus, W., Nitsche, M.A., Pascual-Leone, A.,
using transcranial magnetic stimulation. J Physiol 551, 661– Byblow, W.D., Berardelli, A., Siebner, H.R., Classen, J.,
671. Cohen, L.G. & Rothwell, J.C. 2008. Consensus: motor
Todd, G., Taylor, J.L. & Gandevia, S.C. 2004. Reproducible cortex plasticity protocols. Brain Stimul 1, 164–182.
measurement of voluntary activation of human elbow

 2011 The Authors


140 Acta Physiologica  2011 Scandinavian Physiological Society, doi: 10.1111/j.1748-1716.2011.02271.x

View publication stats

You might also like