You are on page 1of 15

Journal of Neural Engineering

PAPER You may also like


- Wearable multichannel haptic device for
Neurophysiological correlates of force control encoding proprioception in the upper limb
Patrick G Sagastegui Alva, Silvia Muceli, S
improvement induced by sinusoidal vibrotactile Farokh Atashzar et al.

- A microfabricated strain gauge array on


stimulation polymer substrate for tactile
neuroprostheses in rats
M Beygi, S Mutlu and B Güçlü
To cite this article: Carina Marconi Germer et al 2020 J. Neural Eng. 17 016043
- Sensory adaptation to electrical stimulation
of the somatosensory nerves
Emily L Graczyk, Benoit P Delhaye,
Matthew A Schiefer et al.

View the article online for updates and enhancements.

This content was downloaded from IP address 131.188.6.50 on 27/10/2022 at 09:41


J. Neural Eng. 17 (2020) 016043 https://doi.org/10.1088/1741-2552/ab5e08

PAPER

Neurophysiological correlates of force control improvement


induced by sinusoidal vibrotactile stimulation
RECEIVED
24 October 2019
RE VISED
22 November 2019
ACCEP TED FOR PUBLICATION
Carina Marconi Germer1 , Alessandro Del Vecchio2, Francesco Negro3, Dario Farina2 and Leonardo Abdala
2 December 2019 Elias1,4,5
1
PUBLISHED Neural Engineering Research Laboratory, Department of Biomedical Engineering, School of Electrical and Computer Engineering,
31 January 2020 University of Campinas, Campinas, Brazil
2
Neuromechanics & Rehabilitation Technology Group, Faculty of Engineering, Department of Bioengineering, Imperial College
London, London, United Kingdom
3
Department of Clinical and Experimental Sciences, Research Centre for Neuromuscular Function and Adapted Physical Activity ‘Teresa
Camplani’, Università degli Studi di Brescia, Brescia, Italy
4
Center for Biomedical Engineering, University of Campinas, Campinas, Brazil
5
Author to whom any correspondence should be addressed.
E-mail: leoelias@unicamp.br

Keywords: force control, vibrotactile stimulation, motor units, first dorsal interosseous, cutaneous mechanoreceptors

Abstract
Objective. An optimal level of vibrotactile stimulation has been shown to improve sensorimotor
control in healthy and diseased individuals. However, the underlying neurophysiological
mechanisms behind the enhanced motor performance caused by vibrotactile stimulation are yet
to be fully understood. Therefore, here we aim to evaluate the effect of a cutaneous vibration on the
firing behavior of motor units in a condition of improved force steadiness. Approach. Participants
performed a visuomotor task, which consisted of low-intensity isometric contractions of the first
dorsal interosseous (FDI) muscle, while sinusoidal (175 Hz) vibrotactile stimuli with different
intensities were applied to the index finger. High-density surface electromyogram was recorded
from the FDI muscle, and a decomposition algorithm was used to extract the motor unit spike trains.
Additionally, computer simulations were performed using a multiscale neuromuscular model to
provide a potential explanation for the experimental findings. Main results. Experimental outcomes
showed that an optimal level of vibration significantly improved force steadiness (estimated as the
coefficient of variation of force). The decreased force variability was accompanied by a reduction
in the variability of the smoothed cumulative spike train (as an estimation of the neural drive to the
muscle), and the proportion of common inputs to the FDI motor nucleus. However, the interspike
interval variability did not change significantly with the vibration. A mathematical approach,
together with computer simulation results suggested that vibrotactile stimulation would reduce
the variance of the common synaptic input to the motor neuron pool, thereby decreasing the low
frequency fluctuations of the neural drive to the muscle and force steadiness. Significance. Our results
demonstrate that the decreased variability in common input accounts for the enhancement in force
control induced by vibrotactile stimulation.

1. Introduction to the fingertip of participants during a postural


task could reduce the body sway. Mendez-Balbuena
Since the study by Collins and coworkers [1] et al [5] and Trenado et al [6] used an index finger
vibrotactile stimulation has been shown to improve sensorimotor task to show that stochastic vibrotactile
both sensory perception and sensorimotor control stimulation could decrease the mean variation of
in humans. For instance, Priplata et al [2, 3] reported index finger position in healthy participants and
an improved balance control when a bandlimited patients with essential tremor, respectively. More
Gaussian vibratory noise stimulated the foot soles of recently, Germer et al [7] reported that a sinusoidal
young and elderly participants. Similarly, Magalhaes vibrotactile stimulation applied to the index finger
and Kohn [4] showed that vibrotactile noise applied could also improve force steadiness during isometric
© 2020 IOP Publishing Ltd
J. Neural Eng. 17 (2020) 016043 C M Germer et al

contractions of the first dorsal interosseous (FDI) the recording of a small sample of neural elements pre-
muscle, but the improvement was more significant cludes a broader overview of the neurophysiological
for low-intensity contractions. In addition to these mechanisms behind the improved sensorimotor con-
behavioral findings, neurophysiological studies trol. Second, short-term synchronization and ISI CoV
have reported that vibrotactile stimulation increases are only partially correlated to force variability [25, 26]
the power of the electroencephalogram (EEG) since the motor neuron pool is an effective linear fil-
recorded in the somatosensory area [8], and beta- ter that transmits common components to the muscle
band corticomuscular coherence (an estimate of the while cancels out the independent inputs [25]. There-
coupling between cortical and spinal circuits) [9]. fore, both short-term synchronization (based on the
The observations discussed above may relate to analysis of pairs of motor units) and ISI CoV are biased
stochastic resonance in the sensorimotor system. Sto- measures due to their dependence on the total fluctua-
chastic resonance is a counterintuitive phenomenon tions of motor neuron membrane potential, and not
in which the capacity of a nonlinear system to detect/ only on the common fluctuations, which is more rel-
transmit a weak input signal can be improved by the evant to mold the force variability [22, 26]. Third, the
addition of noise with appropriate intensity into the findings of a motor task enhanced by electrical stim-
system [10–12]. Although the occurrence of stochastic ulation cannot be directly translated to a vibration-
resonance in an individual neural element is evident induced enhanced motor task.
(e.g. [13]), the neural basis for this phenomenon is An improved sensorimotor control with vibrotac-
somewhat challenging to grasp when a more complex tile stimulation could be explained by (1) a reduced
system (e.g. the neuromuscular system) is taken into low-frequency fluctuation of the common synap-
consideration. Notwithstanding, previous studies sug- tic input to motor neurons [22, 23, 27], and/or (2) a
gest that vibrotactile stimulation would increase the reduced discharge rate variability of motor units
sensitivity of cutaneous mechanoreceptors, thereby (lower ISI CoV) [24]. Although the variability in dis-
improving sensorimotor integration in both spinal charge is partly influenced by the low-frequency fluc-
and supraspinal neural circuits [2, 4–6, 9, 14]. How- tuations of common input (as mentioned above),
ever, these explanations add little to the neurophysio- we hypothesize that vibrotactile stimulation would
logical knowledge since no neural activity was directly mainly influence the fluctuations of common synap-
recorded during the motor-enhanced tasks. tic input, with a smaller effect on independent input
The ultimate neural code of movement is pro- components. For testing this hypothesis, low-intensity
vided by the spinal motor neurons. This code is then isometric contractions of the FDI muscle (index finger
directly converted into function by the contraction of abduction) with and without stimulation of cutane-
the innervated muscle fibers, which together with the ous mechanoreceptors by high-frequency sinusoidal
motor neurons constitute the motor units [15, 16]. vibration [7] were performed. Additionally, computer
Modern technologies for electromyogram (EMG) simulations were carried out on a multiscale model of
signal recording and processing warrant accurate the neuromuscular system [28, 29] as an aid to better
monitoring of the activity of populations of motor understanding the potential mechanisms underlying
units during isometric or quasi-isometric contractions the experimental findings. Partial results were pre-
[17–20]. The analysis of motor unit spike trains pro- sented as an abstract [30].
vides access to the neural determinants of motor tasks.
For example, motor unit behavior directly influences 2.  Materials and methods
the capacity to maintain a stable force. This is mainly
determined by the low-frequency oscillations in the 2.1. Participants
neural drive to muscles encoded by the ensemble of Eleven subjects (25  ±  6 years, 65.45  ±  7.12 kg, 7 men)
motor unit discharges [21–23]. participated in the study. All subjects were right-
Motor unit population analysis would provide handed, had normal or corrected-to-normal vision,
important insights into the effects of vibrotactile stim- and reported no previous history of neuromuscular
ulation on motor performance. However, this popu- diseases. Participants gave their written informed
lation analysis has not been yet possible. The study consent before the experiments. The procedures were
by Kouzaki et al [24] recorded few motor units (ten) in accordance with the Declaration of Helsinki and
using intramuscular EMG electrodes during a motor were approved by the Imperial College Research Ethics
task enhanced by subthreshold electrical stimulation Committee.
delivered percutaneously to the peripheral nerve of
ankle extensor muscles. They found a decrease in both 2.2.  Force measurement and vibrotactile
motor unit discharge rate variability (estimated as the stimulation
coefficient of variation, CoV, of the motor unit inter- Abduction force of the index finger was measured with
spike intervals, ISI) and motor unit synchronization, a three-axis force transducer (Nano25, ATI Industrial
which was associated with an increase in force steadi- Automation) attached to a mechanical apparatus
ness (i.e. better motor performance). However, a few (figure 1(a)). The force signal was digitized at 2048
aspects of the referred study should be stressed. First, Hz (USB-6225, National Instruments) and lowpass

2
J. Neural Eng. 17 (2020) 016043 C M Germer et al

Figure 1.  Experimental protocol and data analysis. (a) Mechanical apparatus used to record index finger abduction force. Isometric
forces were measured at the distal interphalangeal joint and a linear resonant actuator was placed at the radial surface of the
metacarpophalangeal joint of the index finger (i). A high-density (HD) surface EMG grid with 64 electrodes was placed on the FDI
muscle (ii). (b) Representative recordings of force signals from a single participant in control (CTR, black) and optimal vibration
(OV, grey) conditions at 5% of maximum voluntary contraction (MVC). Data analysis was carried out on data confined between
the dashed lines (i.e. between 23 s and 44 s) to avoid non-stationarities. (c) Representative analysis of the influence of vibrotactile
stimulation on force steadiness (force coefficient of variation, CoV) for two participants. The effects of low-intensity (LV),
intermediate-intensity (IV), and high-intensity (HV) vibration on force CoV were normalized to the no-vibration condition (CTR).
Black circles depict the OV, which was defined as the vibration amplitude that produced the lowest force CoV (relative to CTR) for
each participant at each contraction intensity. Note that the participant on the right panel did not have an OV at 2.5%MVC. (d) Data
analysis performed on HD surface EMG signals. Leftmost signals represent the data from the 64 EMG channels (rows) recorded
during CTR (black) and OV (grey) conditions. The EMG channels were decomposed into motor unit spike trains (middle), so that
each row represents one motor unit and the vertical bars represent the discharge timings. The neural drive to the muscle (rightmost
signals) was estimated as the smoothed cumulative spike train (sCST), which consists of a lowpass filtered (LPF) version (at 2 Hz) of
the cumulative spike trains (see section 2.5.3 for details).

filtered (15 Hz cutoff frequency) with a fourth-order nearly imperceptible low-intensity vibration (0.01G);
Butterworth filter. (2) intermediate-intensity vibration (0.45G); and (3)
Vibrotactile stimulation was performed by a linear high-intensity vibration (1.50G).
resonant actuator (resonant frequency of 175 Hz, C10-
100, Precision Microdrives) placed on the radial sur- 2.3.  Recording of high-density surface EMG
face of the metacarpophalangeal joint (figure 1(a)). A A 13  ×   5 flexible grid of electrodes (4  mm
previous study demonstrated the efficacy of this vibra- interelectrode distance, ELSCH064NM4, OT
tion protocol in improving force control [7], similar to Bioelettronica) was placed on the FDI muscle (figure
the effects showed in other studies with tactile noise [5, 1(a)). Skin preparation was performed with light skin
9, 14]. Three intensities of vibration were applied: (1) abrasion and cleaning with alcohol before placement

3
J. Neural Eng. 17 (2020) 016043 C M Germer et al

of the electrode grid. High-density (HD) surface EMG For each contraction intensity (i.e. 2.5%MVC and
signals were recorded with a multichannel amplifier 5%MVC), the average force CoVs calculated from the
(Quattrocento, OT Bioelettronica) in monopolar three repetitions were compared between the vibra-
mode. Signals were bandpass filtered (10–500 Hz) and tion intensities. In order to normalize the subject-
digitized with 16-bit resolution at 2048 Hz. dependent effect of vibrotactile stimulation on sen-
sorimotor performance [5, 6, 9, 14], we selected an
2.4.  Experimental protocol optimal vibration (OV) for each participant. OV was
Participants were comfortably seated on a chair defined post-experiment as the vibration intensity that
with both arms resting on a table. A custom-made provided the lowest average force CoV (figure 1(c))
apparatus supported the right hand in vertical position from the three repetitions, and was determined for
(figure 1(a)). Velcro® straps were used to immobilize each force intensity [7]. Henceforth, the analyses were
the forearm and wrist. The index finger was aligned carried out on dependent variables measured in the
with the forearm and the thumb was kept in a resting two contraction intensities (2.5% and 5%MVC) and
position at the same height as the index finger. An LCD two vibration conditions (CTR and OV).
monitor was placed ~60 cm in front of the subject at
the eyes level to provide a visual feedback of the force. 2.5.2.  HD surface EMG signals
The target force was showed as a red line at the center of HD surface EMG signals were decomposed into
the monitor screen, and the force feedback signal was motor unit spike trains using a blind source separation
provided as a moving yellow line in which the y -axis algorithm (figure 1(d)) [20, 31, 32]. The pulse to noise
position corresponded to the force component normal ratio (PnR) that estimated the accuracy in discharge
to the sensor surface. identification was set to a minimum of 30 dB [20, 31].
The experiment was conducted in three steps: (1) After the automatic decomposition, the motor unit
maximum voluntary contraction (MVC); (2) famil- spike trains were visually inspected by an experienced
iarization; and (3) the main experimental block. To researcher to correct for misclassified spikes. Motor
estimate the MVC, subjects were asked to contract units that exhibited intermittent firing were discarded
their FDI muscle in a maximal effort for 10 s. The MVC from the analysis.
was defined as the maximum value of force obtained
from three attempts. The familiarization consisted 2.5.3.  Motor unit spike trains
of eight trials (all combinations of target forces and We assessed the mean ISI, as well as the standard
vibration intensities used in the main part of the deviation (SD) and CoV of ISIs. The neural drive to the
experiment—see below) with 15 s duration. The main FDI muscle was estimated as the smoothed cumulative
experiment consisted of steady isometric abductions spike train (sCST, figure 1(d)). The CST of each trial
of the index finger at two contraction intensities (2.5% was estimated as the sum of all motor unit spike trains
and 5%MVC). Each trial of the main experiment was and normalized by the number of motor units. The
also performed in a given vibration condition: with- CST was smoothed using a non-overlapping 400 ms
out vibration (control, CTR) and with the applica- duration Hanning window. The mean and CoV of the
tion of a vibrotactile stimulus at three intensities as sCST were calculated for each condition.
defined above (low, intermediate, and high). In trials The proportion of common input (PCI) to the
where vibration was applied, the stimulus started at motor neuron pool was estimated from the coherence
the beginning of the trial up to its end. Trials lasted 45 s, function [33] in the frequency band between 0.20 Hz
and each condition (contraction intensity and vibra- and 2 Hz (equation (1)). Briefly, for each trial, coher-
tion condition) was repeated three times. The presen- ence functions between two groups with m motor
tation order of each combination of target force and units were computed, where m ranged from one to half
vibration intensity in both the familiarization and the number of decomposed motor unit spike trains:
main experiment blocks were randomized to avoid 2
the influence of learning and adaptation. A 30 s resting |m2 A|
Cf1 ,f2 =
(1) 2
interval was included between trials to avoid fatigue. (mB + m2 A)
where, Cf1 ,f2 is the average coherence in the frequency
2.5.  Data analysis band [ f1 , f2 ], A and B are parameters to be optimized.
2.5.1.  Force signals The parameters A and B in equation (1) are related
Force signals were processed offline in MATLAB to the power of the common synaptic input, and the
(The MathWorks Inc.). The force signal was lowpass squared root of the ratio A/B is an estimate of the PCI
filtered (fourth-order Butterworth digital filter, 15 Hz [33]. We solved the nonlinear least-squares optim­
cutoff frequency) and detrended. The initial 23 s and ization method with the function ‘lsqcurvefit’ of
last 1 s of each force time series (from each trial) were MATLAB. The values of Cf1 ,f2 were estimated using
discarded to avoid non-stationarities (figure 1(b)). the pooled coherence between the two time series that
Force variability was measured as the CoV of the force comprised the concatenation of all possible permuta-
signal. tions of groups of m motor units or up to a maximum

4
J. Neural Eng. 17 (2020) 016043 C M Germer et al

Figure 2.  Schematic diagram of the motor neuron pool model for the two scenarios explored here: (a) with a single common input
(corticomotor drive), and (b) with two common inputs (corticomotor and secondary drives). In addition, in both scenarios each
motor neuron received independent commands (noise). Common presynaptic commands were represented by Gamma point
processes with mean interspike intervals (µ) and order (k ) varying according to the values presented between the brackets.

of 1000 permutations [34]. Coherence was estimated input was adopted to represent the activity from cuta-
using the Welch’s averaged periodogram with non- neous afferents (figure 2(b)). It is worth noting that the
overlapping Hanning windows of 5 s duration (reso- cutaneous pathway to motor neurons is oligosynap-
lution of 0.20 Hz). Only trials with at least six motor tic [41], but for simplicity we assumed that the whole
units were used for the PCI analysis. circuit would result in a single common input to the
motor neurons.
2.6.  Computer simulations Synaptic inputs were represented by a first-order
2.6.1.  Model description kinetic model [42], whose parameters followed those
Computer simulations were carried out using a reported in [28] for excitatory synapses (reversal
multiscale neuromuscular model extensively described potential equal to 70 mV, and time constant equal to
elsewhere [28, 29, 35–38]. Here we will provide only a 0.11 ms). Each synapse from common inputs had a
brief description of the model. The original model was maximum conductance of 600 nS, while the synapses
parameterized to represent the leg muscles, but in the from independent inputs had a maximum conduct-
present study we adapted the model to represent the ance of 6000 nS. The latter was larger than the former
motor neuron pool innervating the FDI muscle. The since each motor neuron had a single independent
motor neuron pool consisted of 120 type-identified input, while common inputs were many (100 for cor-
motor neurons (101 S-type, 17 FR-type, and 2 FF- ticomotor drive and 100 for the secondary common
type [39]). Each motor neuron was represented by a input). The common inputs were simulated using
two-compartment model (soma-dendrite). The soma Gamma stochastic point processes (mean ISI equal to
encompassed the passive properties of the membrane 20 ms), while the independent inputs were realizations
(resistance and capacitance), and active ionic currents of Poisson point processes (mean ISI equal to 50 ms for
yielding action potential (fast sodium and potassium all simulations).
currents) and afterhyperpolarization (slow potassium) Muscle unit force was produced by the impulse
time courses. To speed up the simulations, state response of a second-order critically damped system
variables associated to the dynamics of ionic currents [43]. Twitch amplitudes and contraction times ranged
were simplified using the approach by Destexhe [40]. from 1 mN to 140 mN and 30 ms to 100 ms, respec-
The dendrite was a passive compartment, since we have tively [43, 44]. The FDI muscle force was computed as
no clear evidence of the involvement of active dendritic the sum of the forces produced by each active muscle
currents (e.g. persistent inward currents) during low- unit.
intensity isometric contractions. Morphological and The model was implemented in Java programming
electrophysiological parameters were adopted equal language (Oracle), and the differential equations were
to those reported in [28, 29], with a piecewise linear solved using a fourth order Runge-Kutta method with
interpolation of the reference values (see table 2 in a fixed step (50 µs).
[28] and table 1 in [29]) to represent the population of
motor neurons. 2.6.2.  Simulation protocols
The motor neuron pool was commanded by a Each simulation had 10 s duration and 17 simulations
combination of common and independent presynap- were performed to evaluate the variability between
tic inputs (all located at the dendritic compartment). each simulation run. First, the maximum muscle
Two scenarios were explored: (1) a single common force was estimated as the average contraction
input that represented the corticomotor drive (figure produced by a neural drive of ~30 Hz [39]. Subsequent
2(a)); and (2) two common inputs, where one input simulations were performed to achieve submaximal
represented the corticomotor drive and the secondary muscle force. The commands to the motor neuron

5
J. Neural Eng. 17 (2020) 016043 C M Germer et al

pool were adjusted so that the mean ISI and ISI CoV assumed to be independent on the secondary (cutane-
from the model were both similar to the experimental ous afferent) input, the firing rate variance is given by
counterparts. equation (4) (for simplicity, we ignored the influence
In the first scenario (see description above and fig- of skewness, i.e. the last term of equation (3)). We can
ure 2(a)), we tested the hypothesis that the OV would calculate the theoretical variance using the parameters
directly influence the variability of a single common described in section 2.6.2.
input (corticomotor drive). Therefore, the order Å ã
1000 1 1
(kCM ) of the Gamma point process was systematically σG2 = + +
kCM · µ2CM 6 2kCM 2 · µ2
changed from 4 (control) to 8 and 10 (more regular), Å ã CM
while the mean ISI (µCM ) was set to 20 ms. In the sec- 1000 1 1
+ + + 2 2 . (4)
ond scenario (figure 2(b)), we added a secondary com- kS · µ2S 6 2kS · µS

mon input to evaluate the hypothesis that the pathway
from cutaneous afferents is independent from the
corticomotor drive and would differentially influence 2.8.  Statistical analysis
the variance of the global common input to the motor Statistical analysis was performed using SPSS (IBM).
neuron pool. We evaluated the effect of the second- The significance level adopted in the study was 0.05.
ary common input with the same distribution of the Data are represented as mean  ±  95% confidence
control case (i.e. kS = kCM = 4 ) and for more regular interval. Prior to the regression analysis, a normality
Gamma point process (kS = 10). The reciprocal of the test was performed using the Shapiro-Wilk method.
mean ISI (i.e. the mean rate) of the global common The effect size was calculated as the partial eta-
input (µG) followed equation (2), and we evaluated five squared (ηp2). Large effects were considered when
different combinations [µCM ; µS ]: [20; 0] ms, [25; 100] ηp2 > 0.25, while small effects were considered when
ms, [30, 60] ms, [4040; 40] ms, and [30, 6060; 30] ms. ηp2 < 0.01 [46].
In all combinations, the mean rate (mean ISI) of the For the experimental data, a two-way analysis
global common input was at 50 Hz (20 ms) in order to of variance with repeated measures (two-way RM-
maintain the mean excitability of the motor pool equal ANOVA) and Bonferroni’s post hoc tests were used to
across the simulations. compare the dependent variables between contrac-
tion intensities (2.5% and 5%MVC) and vibration
1 1 1
=
(2) + . conditions (CTR and OV). In order to minimize type I
µG µCM µS
error in our analysis due to the selection of an OV out
of three vibration intensities, we corrected the signifi-
2.6.3.  Simulation data analysis cance level to 0.017 (0.05/3).
Simulation data were analyzed in the same way as the Computer simulation data were analyzed using a
experimental data (see section 2.5). The dependent multiple ANOVA and Dunnet’s tests to compare the
variables measured from the simulations were: mean dependent variables measured in different input con-
force; force CoV; mean ISI from the motor neurons; ditions with those in the control condition (kCM = 4
motor neuron ISI CoV; mean sCST; sCST CoV; PCI; and µCM = 20 ms). In the scenario with a single com-
and the resultant variance of the common input mon input (figure 2(a)), the control condition was
(estimated from a smoothed version of the spike compared with two conditions where the order of
trains composed of the Gamma point processes that the corticomotor input was changed (kCM = 8 and
represent the common inputs to the motor neuron 10 ). Similarly, in the scenario with two common inputs
pool). (figure 2(b)) the control condition was compared with
different conditions where the intensities of the two
2.7.  Theoretical calculation of common input inputs were systematically varied (µCM /µS equal to
variance 25/100, 30/60, 40/40, and 60/30).
From the theory of stochastic point processes, the
variance of ISI of a Gamma point process (renewal 3. Results
2
process) is given by σISI = µ/k, where µ is the mean
ISI (in milliseconds) and k is the order of the Gamma 3.1.  Human experiments
point process. Similarly, the asymptotic variance of the All participants were able to maintain the muscle force
2
firing rate (σFR ) is given by equation (3) [45]. at the target levels regardless of the vibration intensity
2 ( p = 0.758 ). In total, 295 (305) and 357 (350) motor
2 σ 2 · 1, 000 1 (σISI
2
) µ3
(3)
σFR = ISI 3 + + 4
− 3 units were decomposed in CTR (OV) at 2.5% and
µ 6 2µ 3µ 5%MVC, respectively. The following analyses were
where, µ3 is the skewness of the ISI distribution. performed with the spike trains from these motor
In this section of the present study, we aimed to cal- units.
culate the variance of the global common input to the No interaction between force intensities (2.5% and
motor neuron pool. Since the corticomotor input is 5%MVC) and vibration conditions (CTR and OV)

6
J. Neural Eng. 17 (2020) 016043 C M Germer et al

Table 1.  Outcomes from the two-way repeated measures (RM) ANOVA performed on the experimental data. Asterisks and bold numbers
indicate the significant differences ( p < 0.017 ). Effect sizes are reported as the partial eta-squared (ηp2 ).

Mean
Mean Force Force CoV Mean ISI ISI CoV sCST sCST CoV PCI

Force p <0.001* <0.001* 0.003* 0.312 0.004* 0.087 0.743


F(1, 10) 37 944.746 75.691 14.967 1.133 13.616 3.603 0.115
ηp2 1 0.883 0.599 0.102 0.577 0.265 0.013

Vibration p 0.964 <0.001* 0.354 0.148 0.465 0.012* 0.004*


F(1, 10) 0.002 28.345 0.945 2.465 0.577 9.508 14.225
ηp2 0.000 0.739 0.086 0.198 0.055 0.487 0.612

Interaction p 0.591 0.070 0.271 0.985 0.199 0.216 0.069


(force  ×  vibration) F(1, 10) 0.308 4.115 1.358 0.000 2.909 1.741 4.279
ηp2 0.030 0.292 0.120 0.000 0.225 0.148 0.322

Figure 3.  Experimental results. (a) Force coefficient of variation (CoV) as a function of force intensities (2.5% and 5%MVC) and
vibration conditions (control, CTR, and optimal vibration, OV). (b)–(d) The same as (a), but for the interspike (ISI) CoV, sCST,
and proportion of common input (PCI), respectively. Error bars represent the 95% confidence interval (n = 11). Asterisks indicate
significant differences for vibration conditions ( p < 0.017 ).

was observed for the dependent variables measured spective of the force intensity. Also, OV had no effect
from force and from the motor unit spike trains (i.e. on mean ISI, mean sCST, and ISI CoV (figure 3(b))
force CoV, ISI CoV, sCST CoV, and PCI; see table 1). but significantly decreased sCST CoV and PCI (fig-
Increasing the intensity of contraction from ures 3(c) and (d)).
2.5%MVC to 5%MVC was followed by a significant
increase in mean sCST and a decrease in force CoV 3.2.  Computer simulations
and mean ISI. Conversely, no significant difference was In the first scenario simulated in this study (see
found for ISI CoV, sCST CoV and PCI. section 2.6.2 for details), a more regular corticomotor
When the target force was 5%MVC at least one input (the only common input impinging onto the
vibration intensity was able to decrease the force CoV motor neurons) significantly decreased all dependent
(improvement in steadiness) for all participants. variables as compared to the control condition
However, at 2.5%MVC no OV was found for two par- (kCM = 4) (figure 4), without changing the mean force
ticipants of the study. Figure 3(a) and table 1 show (table 2). These results contrast to the experimental
that force CoV was significantly decreased (large outcomes since in the experiments we did not observe
effect) in OV as compared to the CTR condition, irre- significant changes in the ISI CoV (see figure 3(b)).

7
J. Neural Eng. 17 (2020) 016043 C M Germer et al

Figure 4.  Computer simulation results for the scenario where the variability of the corticomotor input (single common input) was
reduced (kCM = 8 and kCM = 10) with respect to the control condition (kCM = 4). Effects on (a) force CoV, (b) ISI CoV, (c) sCST
CoV, and (d) PCI. Error bars represent the 95% confidence interval (n = 17). Asterisks indicate significant differences ( p < 0.05).

Table 2.  Outcomes from the multiple ANOVA performed on the computer simulation data. Asterisks and bold numbers indicate
significant differences ( p < 0.05). Effect size is reported as the partial eta-squared (ηp2 ).

Mean Mean Mean


force Force CoV ISI ISI CoV sCST sCST CoV PCI Variance

µCM = 20 Order p 0.517   <  0.001* 0.240   <  0.001* 0.115   <  0.001* 0.001*   <  0.001*
(kCM ) F(2, 62) 0.666 60.566 1.462 25.946 2.239 33.045 7.765 28.522
ηp2 0.21 0.661 0.0445 0.456 0.067 0.516 0.200 0.479
Multiple 4 versus 8 0.225   <  0.001* 0.745   <  0.001* 0.807   <  0.001* 0.033*   <  0.001*
compariso­ns 4 versus 10 0.615   <  0.001* 0.115   <  0.001* 0.998   <  0.001*   <  0.001*   <  0.001*

kS = 4 Condition p 0.097 0.490 0.651 0.049* 0.268 0.307 0.453 0.897


F(3, 64) 2.199 0.815 0.548 2.765 1.343 1.227 0.887 0.199
ηp2 0.093 0.037 0.025 0.115 0.059 0.054 0.040 0.009
Multiple versus 25 0.726 0.832 0.956 0.998 0.145 0.361 0.777 0.480
compariso­ns versus 30 0.997 0.867 0.966 1.000 0.098 0.325 0.881 0.596
(µCM = 20 ) versus 40 0.996 0.990 0.977 0.995 0.142 0.073 0.285 0.754
versus 60 — — — — — — — —

kS = 10 Condition p 0.237 0.003* 0.955 0.014* 0.907   <  0.001* 0.006*   <  0.001*


F(4, 80) 1.414 4.439 0.167 3.331 0.253 7.217 3.942 9.203
ηp2 0.660 0.182 0.008 0.143 0.013 0.265 0.165 0.315
Multiple versus 25 0.988 0.885 0.901 0.968 0.469 0.066 0.858 0.055
compariso­ns versus 30 0.999 0.666 0.934 0.998 0.473 0.026* 0.657 <0.001
(µCM = 20) versus 40 0.968 0.012 * 0.798 0.316 0.699   <  0.001 *
0.014 *
  <  0.001*
* * *
versus 60 0.794 0.018 0.772 0.292 0.722   <  0.001 0.028   <  0.001*

The addition of a secondary input (second sce- ular input) there was a significant decrease in force
nario described in section 2.6.2) with kS = 4 did not CoV, sCST CoV, and PCI when the intensity of the
significantly change any dependent variable evalu- secondary input increased (µS = 40 ms and µS = 30
ated in the study (see figures 5(a)–(d)). Conversely, ms). Also, in the latter condition, no difference was
figures 5(e)–(h) shows that when kS = 10 (more reg- observed for the mean force, mean ISI, ISI CoV, and

8
J. Neural Eng. 17 (2020) 016043 C M Germer et al

Figure 5.  Computer simulation results for the scenario where the secondary common input was added to the model. (a)–(d) Results
achieved when kCM = kS = 4 . (e)–(h) Results achieved when kCM = 4 and kS = 10 (more regular input). Effects are shown for: (a)
and (e) force CoV; (b) and (f) ISI CoV; (c) and (g) sCST CoV; and (d) and (h) PCI. Error bars represent the 95% confidence interval
(n = 17). Asterisks indicate significant differences ( p < 0.05).

Figure 6.  Theoretical analysis on the effects of order (k ) and mean intensity (µ) of the two common inputs (Gamma point
processes) on the variance of the global common input (σG2 ). Panels (a)–(d) show analytical curves for different orders of the
corticomotor input (i.e. kCM = 1, 4, 6, and 10 , respectively).

mean sCST among the different combinations of com- 3.3.  Variance of the common input
mon input intensities, which is compatible with the Figure 6 illustrates the effect of a secondary input on
experimental results presented in section 3.1). Table 2 the resultant variance of the global common input.
shows a complete description of the statistics from the These results are based on the theoretical approach
computer simulation data. described in section 2.7. If the corticomotor input

9
J. Neural Eng. 17 (2020) 016043 C M Germer et al

was represented as a stochastic Poisson point process Therefore, the experimental findings can be explained
(i.e. a stochastic Gamma point process with order by a reduction in the variance of the global common
kCM = 1), the addition of a secondary input of any input due to the activation of a secondary input, which
order (kS) decreased the resultant input variance we hypothesize to be the neural activity induced by the
(figure 6(a)). Conversely, when the regularity vibrotactile stimulation. In the following subsections,
of the corticomotor input increased (i.e. higher we shall present a detailed discussion on each of these
kCM ) the input variance depended on the regularity findings.
(kS) and intensity (µS ) of the secondary input. When
kCM = 4, the common input variance decreased only 4.1.  Effects of an optimal vibrotactile stimulation
when kS  5, irrespective of the µS (figure 7(b)). on a population of motor units
However, when kS = 3 and kS = 4 , there were optimal A relevant experimental finding was that OV did not
regions where the balance between the intensities of reduce either ISI CoV or the relationship between ISI
the corticomotor and secondary inputs produced a SD and the mean ISI (data not shown), but reduced
minimum input variance. For kCM = 6 (figure 6(c)) the fluctuations of the neural drive to the muscle
a decrease in the global common input variance was (estimated as the sCST [22]), as well as the relative
observed only when kS  7 , but the observed decrease proportion of common input (figure 3 and table 1).
was almost negligible. Finally, when kCM = 10 the Although the motor neuron activity is intrinsically
resultant global common input variance increased for nonlinear, the population of recruited motor units acts
any intensity and order of the secondary input (figure as a lowpass linear system [22, 25, 47]. In other words,
6(d)). the population of motor units linearly transmits the
Figure 7 shows a comparison between computer slowly varying components of the common input to
simulation data and the theoretical analysis of the the motor output and filters out the independent syn-
global common input variance. Figure 7(a) shows aptic inputs received by each motor neuron [23]. Our
that when the common input is represented by a sin- findings support this idea since the reduction in force
gle corticomotor drive the input variance significantly CoV caused by OV was not followed by a reduction in
decreased (as expected from the theoretical curve; ISI CoV but by a reduction in sCST CoV, which repre-
see dashed line) when kCM increased (more regular). sents the variability of low-frequency components (up
When the two common inputs are considered in the to 2.50 Hz) of the common input. Despite the evidence
model, there was no significant difference in the input of correlation between ISI CoV and force CoV [24, 48],
variance when kCM = kS = 4 (figure 7(b)), whereas the current opinion is that the motor output variability
significant differences were observed when kS = 10 is mainly influenced by the low-frequency fluctuations
and µs < 60 ms (figure 7(c)). These differences in of the common inputs to the motor neuron pool [22,
the common input variance are related to the changes 49, 50].
observed in force CoV, sCST CoV, and PCI from comp­ The PCI is an estimate of the relative propor-
uter simulation data. tion between the common inputs impinging onto
the motor neurons and the total synaptic inputs at
4. Discussion low frequencies [33]. This index may also be viewed
as the proportion between the common fluctuations
We combined human experiments and computer on motor neuron membrane potential and the total
simulations to provide clues on the neurophysiological membrane potential fluctuations (sum of common
mechanisms behind the improved force control and independent inputs). Therefore, a decreased PCI
induced by sinusoidal vibrotactile stimulation. could be a consequence of a reduced fluctuation of the
In the experiments, we observed that an optimal common inputs and/or an increased fluctuation of the
level of vibrotactile stimulation (OV) could reduce independent inputs. Another possible explanation for
force variability (CoV) accompanied by a decrease the decreased PCI is a decorrelation between the activ-
in the variability of the neural drive to the FDI ities of the motor units caused by activation of cuta-
muscle (reduced sCST CoV) and a reduction in the neous afferents. Kouzaki et al [24] showed that motor
proportion of common inputs (PCI) to the recruited unit synchronization was reduced when a subthresh-
motor units. However, OV did not change ISI old stochastic electrical stimulation was applied to the
variability as compared to the control (no vibration) tibial nerve. Despite the expected difference between
condition. Computer simulations could resemble the the experimental protocol of the referred study and
experimental findings only when a combination of ours (i.e. their protocol preferably excited muscle
two independent common inputs were used in the afferents), the reduced motor unit synchronization is
model, and the secondary input had a lower variability. in the same direction of our finding of reduced PCI
Additionally, from the theory of stochastic point (although we did not find any difference in motor
processes, we showed that the variance of the global unit synchronization—data not shown). Since PCI is
common input could be reduced by an appropriate a measure derived from the coherence between two
balance between intensity and variability of the two CSTs, it also reflects the degree of correlation imposed
independent inputs impinging on the motor pool. by common inputs to the motor neuron pool [33].

10
J. Neural Eng. 17 (2020) 016043 C M Germer et al

Figure 7.  Variance of the global common input. (a) Relationship between the common input variance (σG2 ) and the order of
the corticomotor input (kCM ), which was the single input in the model for this scenario. (b) and (c) Input variance when two
common inputs are considered in the model. Panel (b) shows the relationship between the input variance and the intensity of the
corticomotor input (µCM ) when kCM = kS = 4 , while panel (c) shows the same relationship when kS = 10. Error bars represent the
95% confidence interval (n = 17). Asterisks indicate significant differences ( p < 0.05).

Therefore, the activation of cutaneous afferents by equation (4)). With a more regular secondary input,
vibrotactile stimulation would reduce the corticomo- the global common input will have a lower variability
tor drive (but the net excitatory input is maintained, that implies a more regular neural drive to the mus-
since we did not observe changes in the mean excit- cle (sCST with lower variability), and consequently a
ability of the motor units in OV as compared to CTR), steadier force [25, 26, 53, 54]. Second, the activation
thereby decreasing the correlated discharge of motor of the secondary input would decorrelate the cortico-
units [24, 51]. motor input and motor neuron pool output, similar
to the findings by Negro and Farina [51]. Both factors
4.2.  On the influence of common input statistics (i.e. reduced global input variance and decorrelation
A theoretical analysis and computer simulations of corticomotor drive) lead to a reduction of the PCI.
showed that the statistical properties of common When individual motor units were evaluated, as
inputs could largely influence the motor output mentioned before, we did not find modifications in ISI
variability, which was similar to the findings of variability between CTR and OV conditions (figure 3).
Watanabe et al [29]. When a single common input was A simple decrease in the common input variance will
considered in the model (to represent the corticomotor lead to a decrease in ISI CoV, as observed in the simu-
drive), the reduced variability of the presynaptic lations with a single common input (figure 4). The
command reduced force CoV, ISI CoV, sCST CoV, addition of a second input is not expected to change
and PCI (see figure 4). The latter results contrast with this property, but a reduction in the ISI CoV did not
the experimental findings and, therefore, suggest that accompany the lower variance of the global input
vibrotactile stimulation would not directly influence when the second input was added. It is well known
the variance of the corticomotor drive, which is that motor neuron ISI variability depends largely on
compatible with experimental data from monkeys both afterhyperpolarization time course, as well as the
[52]. When the intensity of the global input was low-frequency and high-frequency fluctuations of the
adjusted to produce simulated motor unit discharges motoneuron’s membrane potential [55, 56]. The for-
at the same mean rate of the experimental data, the mer is an intrinsic property that is implausible to be
best scenario to reproduce the experimental outcomes influenced by OV. Therefore, a possible explanation for
was with a reduced variability of the second input as the unchanged ISI CoV is that the nonlinearity in the
compared to the variability of the corticomotor drive. spiking generation process produces constructive and
In this scenario, the activity of the secondary input destructive interferences when two common inputs
(along with the corticomotor drive) reduced the force concurrently activate the motor neuron [51]. This
CoV, sCST CoV, and PCI, but did not change the ISI interference induced by the two inputs limits the trans-
CoV (figure 5 and table 2). mission capacity of any common input to the output
The addition of a secondary input would influ- of a single motor neuron, and hence the independent
ence the activity of motor units in two different (but synaptic noise will play a significant role in shaping the
complementary) ways. First, the analytical approach variability of the motor neuron ISIs. Since we did not
(figures 6 and 7) demonstrated that the variance of change the independent noise across simulated condi-
the global common input (a combination of cortico- tions, the ISI CoV did not change accordingly.
motor and secondary inputs) is mostly influenced by Another insight provided by the theoretical analysis
both the intensity and variability of each input (see and computer simulations is on the existence of condi-

11
J. Neural Eng. 17 (2020) 016043 C M Germer et al

tions with increased common input variance depend- metacarpophalangeal joint. However, there is an
ing on the combination of regularity and intensity of extensive literature showing that tendon vibration
the two common inputs. In our study and several pre- would influence the activity of motor units differently
vious ones [5, 7, 9, 14], the beneficial effect of vibrotac- from the findings reported here [63–66]. For instance,
tile stimulation was not observed for some intensity of tendon vibration increases motor unit firing rate [64],
a vibratory stimulus. For a single subject, some intensi- decreases discharge variability [67], and changes the
ties worsen force control (i.e. increase force CoV, see recruitment strategy of motor units [68]. Since we did
figure 1(c) for an example). Also, the vibration inten- not observe most of these effects in our data (nota-
sity for optimal performance is frequently different bly, the increase in firing rate and ISI variability), we
across the subjects [9]. A possible explanation is that suggest that the direct activation of Ia afferents by
the global common input variance depends on a bal- our vibrotactile stimulation is minimal, and another
ance between the regularities and intensities of the two neurophysiological mechanism is behind the findings
inputs (figure 6). Small modifications in both variabil- reported elsewhere (e.g. tonic vibration reflex).
ity and/or intensity of the corticomotor drive or neural Our computer simulation results suggest that
activity induced by the vibrotactile stimulation would vibrotactile stimulation would activate a more regular
produce an input variance that is higher (or equal) (less variable) neural control pathway during an OV
than (to) a control condition, thereby decreasing (or condition. This hypothesis is plausible since the Pacin-
unchanging) the motor performance. ian corpuscles (thought to be activated by our stimula-
tion protocol [69]) have a preferred firing range, which
4.3.  On the role of cutaneous afferents on motor depends on the vibration intensity and frequency
unit activity and force enhancement [70]. Spike train histograms from the Pacinian affer-
Hitherto, our focus was on the influence of optimal ent activities exhibited a marked periodicity with small
vibrotactile stimulation on the activity of motor units. deviations from the preferred firing frequency. This
However, other experimental studies using different feature is still present even when a bandlimited (100
protocols have shown that stimulation (either Hz–300 Hz) stochastic input is applied to the mecha-
electrical or mechanical) of cutaneous afferents can (i) noreceptor [70]. Therefore, our results and interpre-
reduce the presynaptic inhibition of Ia muscle spindle tations can also be useful to explain the experimental
afferents [57–59], (ii) induce differential excitatory outcomes from studies that used a stochastic vibrotac-
and inhibitory effects on high-threshold and low- tile stimulation to improve force control [5, 9, 14].
threshold motor units, respectively [60], and (iii)
influence the recruitment pattern of motor units by
Acknowledgments
changing their recruitment thresholds [61]. Since we
evaluated the activity of a population of motor units
CMG was the recipient of a PhD scholarship from
rather than tracking individual motor units between
CAPES. She also received a Visiting Student Grant
vibration and no-vibration conditions, we cannot
from PDSE/CAPES (CAPES Foundation, Ministry
rule out the possibility that cutaneous afferents would
of Education, Brazil, proc. no. 88881.134842/2016-
induce differential excitatory/inhibitory effects on
01). LAE is currently funded by CNPq (Brazilian NSF,
the motor units. However, a different experimental
proc. no. 312442/2017-3) and FAPESP (The Sao Paulo
protocol is necessary to explore these aspects of our
Research Foundation, proc. no. 2017/22191-3).
vibrotactile stimulation protocol.
Changes in presynaptic inhibition can also be a
putative neurophysiological mechanism underly- ORCID iDs
ing the findings of the present study. Aimonetti et al
[59] reported that cutaneous vibration could reduce Carina Marconi Germer https://orcid.org/0000-
presynaptic inhibition of Ia terminals without chang- 0002-7323-3767
ing the mean discharge rate of the motor units. A Dario Farina https://orcid.org/0000-0002-7883-
recent computer simulation study showed that force 2697
CoV, common drive index, and motor unit coherence Leonardo Abdala Elias https://orcid.org/0000-0003-
decreased with an increased gain of the monosynap- 4488-3063
tic Ia pathway [62]. In this vein, the Ia afferent activity
would also comprise the hypothetical secondary input
of our model, and its effects would be quite similar to References
those described in the previous sections. Nonetheless, [1] Collins J J, Imhoff T T and Grigg P 1996 Noise-enhanced
the relative contribution of Ia and cutaneous to the tactile sensation Nature 383 770
finding should be evaluated in another protocol mix- [2] Priplata A, Niemi J, Salen M, Harry J, Lipsitz L A and Collins J J
ing vibrotactile and electrotactile stimulations. 2002 Noise-enhanced human balance control Phys. Rev. Lett.
89 238101
Additionally, even in our protocol, we cannot [3] Priplata A A, Niemi J B, Harry J D, Lipsitz L A and Collins J J
rule out a direct activation of Ia afferents due to the 2003 Vibrating insoles and balance control in elderly people
proximity between the vibrotactile actuator and the Lancet 362 1123–4

12
J. Neural Eng. 17 (2020) 016043 C M Germer et al

[4] Magalhaes F H and Kohn A F 2011 Vibratory noise to the [24] Kouzaki M, Kimura T, Yoshitake Y, Hayashi T and Moritani T
fingertip enhances balance improvement associated with light 2012 Subthreshold electrical stimulation reduces motor unit
touch Exp. Brain Res. 209 139–51 discharge variability and decreases the force fluctuations of
[5] Mendez-Balbuena I, Manjarrez E, Schulte-Monting J, plantar flexion Neurosci. Lett. 513 146–50
Huethe F, Tapia J A, Hepp-Reymond M-C and Kristeva R 2012 [25] Dideriksen J L, Negro F, Enoka R M and Farina D 2012 Motor
Improved sensorimotor performance via stochastic resonance unit recruitment strategies and muscle properties determine
J. Neurosci. 32 12612–8 the influence of synaptic noise on force steadiness
[6] Trenado C, Amtage F, Huethe F, Schulte-Mönting J, Mendez- J. Neurophysiol. 107 3357–69
Balbuena I, Baker S N, Baker M, Hepp-Reymond M-C, [26] Farina D and Negro F 2015 Common synaptic input to motor
Manjarrez E and Kristeva R 2014 Suppression of enhanced neurons, motor unit synchronization, and force control Exerc.
physiological tremor via stochastic noise: initial observations Sport Sci. Rev. 43 23–33
PLoS One 9 e112782 [27] Farina D, Holobar A, Merletti R and Enoka R M 2010
[7] Germer C M, Moreira L S and Elias L A 2019 Sinusoidal Decoding the neural drive to muscles from the surface
vibrotactile stimulation differentially improves force electromyogram Clin. Neurophysiol. 121 1616–23
steadiness depending on contraction intensity Med. Biol. Eng. [28] Cisi R R L and Kohn A F 2008 Simulation system of spinal cord
Comput. 57 1813–22 motor nuclei and associated nerves and muscles, in a Web-
[8] Manjarrez E, Diez-Martı́nez O, Méndez I and Flores A 2002 based architecture J. Comput. Neurosci. 25 520–42
Stochastic resonance in human electroencephalographic [29] Watanabe R N et al 2013 Influences of premotoneuronal
activity elicited by mechanical tactile stimuli Neurosci. Lett. command statistics on the scaling of motor output variability
324 213–6 during isometric plantar flexion J. Neurophysiol. 110 2592–606
[9] Trenado C, Mendez-Balbuena I, Manjarrez E, Huethe F, [30] Germer C M, Del Vecchio A, Negro F, Elias L A and Farina D
Schulte-Mönting J, Feige B, Hepp-Reymond M-C and 2018 Effects of vibrotactile stimulation on force steadiness
Kristeva R 2014 Enhanced corticomuscular coherence by on the behavior of motor units of the first dorsal interosseous
external stochastic noise Frontiers Hum. Neurosci. 8 1–10 muscle XXII Int. Society of Electrophysiology and Kinesiology
[10] Moss F 2004 Stochastic resonance and sensory information (Dublin, Ireland)
processing: a tutorial and review of application Clin. [31] Holobar A and Zazula D 2007 Multichannel blind source
Neurophysiol. 115 267–81 separation using convolution Kernel compensation IEEE
[11] McDonnell M D and Ward L M 2011 The benefits of noise Trans. Signal Process. 55 4487–96
in neural systems: bridging theory and experiment Nat. Rev. [32] Negro F, Muceli S, Castronovo A M, Holobar A and Farina D
Neurosci. 12 415–26 2016 Multi-channel intramuscular and surface EMG
[12] McDonnell M D and Abbott D 2009 What is stochastic decomposition by convolutive blind source separation J.
resonance? Definitions, misconceptions, debates, and its Neural Eng. 13 026027
relevance to biology PLoS Comput. Biol. 5 e1000348 [33] Negro F, Yavuz U Ş and Farina D 2016 The human motor
[13] Cordo P, Inglis J T, Verschueren S, Collins J J, Merfeld D M, neuron pools receive a dominant slow-varying common
Rosenblum S, Buckley S and Moss F 1996 Noise in human synaptic input J. Physiol. 594 5491–505
muscle spindles Nature 383 769–70 [34] Amjad A M, Halliday D M, Rosenberg J R and Conway B A
[14] Trenado C, Mikulić A, Manjarrez E, Mendez-Balbuena I, 1997 An extended difference of coherence test for comparing
Schulte-Mönting J, Huethe F, Hepp-Reymond M-C and and combining several independent coherence estimates:
Kristeva R 2014 Broad-band Gaussian noise is most effective in theory and application to the study of motor units and
improving motor performance and is most pleasant Frontiers physiological tremor J. Neurosci. Methods 73 69–79
Hum. Neurosci. 8 22 [35] Elias L A, Chaud V M and Kohn A F 2012 Models of passive
[15] Sherrington C S 1925 Remarks on some aspects of reflex and active dendrite motoneuron pools and their differences in
inhibition Proc. R. Soc. B 97 519–45 muscle force control J. Comput. Neurosci. 33 515–31
[16] Heckman C J and Enoka R M 2012 Motor unit Compr. Physiol. [36] Elias L A and Kohn A F 2013 Individual and collective
2 2629–82 properties of computationally efficient motoneuron models of
[17] Martinez-Valdes E, Laine C M, Falla D, Mayer F and Farina D types S and F with active dendrites Neurocomputing 99 521–33
2016 High-density surface electromyography provides [37] Watanabe R N and Kohn A F 2015 Fast oscillatory commands
reliable estimates of motor unit behavior Clin. Neurophysiol. from the motor cortex can be decoded by the spinal cord for
127 2534–41 force control J. Neurosci. 35 13687–97
[18] Farina D and Holobar A 2016 Characterization of human [38] Watanabe R and Kohn A 2017 Nonlinear frequency-domain
motor units from surface EMG decomposition Proc. IEEE analysis of the transformation of cortical inputs by a
104 353–73 motoneuron pool-muscle complex IEEE Trans. Neural Syst.
[19] Farina D, Negro F, Muceli S and Enoka R M 2016 Principles Rehabil. Eng. 25 1930–39
of motor unit physiology evolve with advances in technology [39] Enoka R M and Fuglevand A J 2001 Motor unit physiology:
Physiology 31 83–94 some unresolved issues Muscle Nerve 24 4–17
[20] Holobar A, Minetto M A and Farina D 2014 Accurate [40] Destexhe A 1997 Conductance-based integrate-and-fire
identification of motor unit discharge patterns from high- models Neural Comput. 9 503–14
density surface EMG and validation with a novel signal-based [41] Pierrot-Deseilligny E and Burke D 2012 The Circuitry of the
performance metric J. Neural Eng. 11 016008 Human Spinal Cord: Spinal and Corticospinal Mechanisms of
[21] Feeney D F, Mani D and Enoka R M 2018 Variability in Movement (Cambridge: Cambridge University Press)
common synaptic input to motor neurons modulates both (https://doi.org/10.1017/CBO9781139026727)
force steadiness and pegboard time in young and older adults [42] Destexhe A, Mainen Z F and Sejnowski T J 1994 An efficient
J. Physiol. 16 3793–806 method for computing synaptic conductances based on a
[22] Negro F, Holobar A and Farina D 2009 Fluctuations in kinetic-model of receptor-binding Neural Comput. 6 14–8
isometric muscle force can be described by one linear [43] Milner-Brown H S, Stein R B and Yemm R 1973 The contractile
projection of low-frequency components of motor unit properties of human motor units during voluntary isometric
discharge rates J. Physiol. 587 5925–38 contractions J. Physiol. 228 285–306
[23] Farina D, Negro F and Dideriksen J L 2014 The effective neural [44] Fuglevand A J, Winter D A and Patla A E 1993 Models of
drive to muscles is the common synaptic input to motor recruitment and rate coding organization in motor-unit pools
neurons J. Physiol. 592 3427–41 J. Neurophysiol. 70 2470–88

13
J. Neural Eng. 17 (2020) 016043 C M Germer et al

[45] Cox D R 1962 Renewal Theory (London: Butle & Tanner) [59] Aimonetti J M, Vedel J P, Schmied A and Pagni S 2000
[46] Cohen J 1988 Statistical Power Analysis for the Behavioral Mechanical cutaneous stimulation alters Ia presynaptic
Sciences (Hillsdale, NJ: Lawrence Erlbaum Associates) inhibition in human wrist extensor muscles: a single motor
[47] Negro F and Farina D 2011 Linear transmission of cortical unit study J. Physiol. 522 137–45
oscillations to the neural drive to muscles is mediated by [60] Datta A and Stephens J A 1981 The effects of digital nerve
common projections to populations of motoneurons in stimulation on the firing of motor units in human first dorsal
humans J. Physiol. 589 629–37 interosseous muscle J. Physiol. 318 501–10
[48] Moritz C T, Barry B K, Pascoe M A and Enoka R M 2005 [61] Garnett R and Stephens J A 1981 Changes in the recruitment
Discharge rate variability influences the variation in force threshold of motor units produced by cutaneous stimulation
fluctuations across the working range of a hand muscle J. in man J. Physiol. 311 463–73
Neurophysiol. 93 2449–59 [62] Nagamori A, Laine C M and Valero-Cuevas F J 2018 Cardinal
[49] Moon H, Kim C, Kwon M, Chen Y T, Onushko T, Lodha N and features of involuntary force variability can arise from the
Christou E A 2014 Force control is related to low-frequency closed-loop control of viscoelastic afferented muscles PLoS
oscillations in force and surface EMG PLoS One 9 e109202 Comput. Biol. 14 e1005884
[50] Lodha N and Christou E A 2017 Low-frequency oscillations [63] Tenan M S, Tweedell A J, Haynes C A and Passaro A D 2019
and control of the motor output Frontiers Physiol. 8 1–9 The effect of imperceptible Gaussian tendon vibration on the
[51] Negro F and Farina D 2011 Decorrelation of cortical inputs Hoffmann reflex Neurosci. Lett. 706 123–7
and motoneuron output J. Neurophysiol. 106 2688–97 [64] Kiehn O and Eken T 1997 Prolonged firing in motor units:
[52] Baker S N, Chiu M and Fetz E E 2006 Afferent encoding of central evidence of plateau potentials in human motoneurons? J.
oscillations in the monkey arm J. Neurophysiol. 95 3904–10 Neurophysiol. 78 3061–8
[53] Del Vecchio A, Ubeda A, Sartori M, Azorin J M, Felici F and [65] Mosier E M, Herda T J, Trevino M A and Miller J D 2017 The
Farina D 2018 The central nervous system modulates the influence of prolonged vibration on motor unit behavior
neuromechanical delay in a broad range for the control of Muscle and Nerve 55 500–7
muscle force J. Appl. Physiol. 125 1404–10 [66] Barrera Curiel A, Colquhoun R J, Hernandez-Sarabia J and
[54] Thompson C K, Negro F, Johnson M D, Holmes M R, DeFreitas J M 2019 The effects of vibration-induced altered
McPherson L M, Powers R K, Farina D and Heckman C J 2018 stretch reflex sensitivity on maximal motor unit firing
Robust and accurate decoding of motoneuron behaviour and properties J. Neurophysiol. 121 2215–21
prediction of the resulting force output J. Physiol. 596 2643–59 [67] Harwood B, Cornett K M D D, Edwards D L, Brown R E and
[55] Calvin W H and Stevens C F 1967 Synaptic noise as a source Jakobi J M 2014 The effect of tendon vibration on motor
of variability in the interval between action potentials Science unit activity, intermuscular coherence and force steadiness
155 842–4 in the elbow flexors of males and females Acta Physiol.
[56] Matthews P B C 1996 Relationship of firing intervals of 211 597–608
human motor units to the trajectory of post-spike after- [68] Xu L, Negro F, Xu Y, Rabotti C, Schep G, Farina D and Mischi M
hyperpolarization and synaptic noise J. Physiol. 492 597–628 2018 Does vibration superimposed on low-level isometric
[57] Nakashima K, Rothwell J C, Day B L, Thompson P D and contraction alter motor unit recruitment strategy? J. Neural
Marsden C D 1990 Cutaneous effects on presynaptic Eng. 15 066001
inhibition of flexor Ia afferents in the human forearm. J. [69] Abraira V E and Ginty D D 2013 The sensory neurons of touch
Physiol. 426 369–80 Neuron 79 618–39
[58] Iles J F 1996 Evidence for cutaneous and corticospinal [70] Bolanowski S J and Zwislocki J J 1984 Intensity and frequency
modulation of presynaptic inhibition of Ia afferents from the characteristics of pacinian corpuscles. I. Action potentials. J.
human lower limb J. Physiol. 491 197–207 Neurophysiol. 51 793–811

14

You might also like