ArticlePDF Available

Quantifying sources of variability in gait analysis

Authors:

Abstract and Figures

Measurements from gait analysis are affected by many sources of variability. Schwartz et al. [1] illustrated an experimental design and methods to estimate these variance components. However, the derivation contains errors which could severely bias the estimation of some components. Therefore, in this paper, we presented correction to this method using ANOVA and Likelihood methods. Furthermore, we demonstrated how commonly used reliability indices like CMC and ICC may be derived from the variance components. We advocate the use of the variance components, in preference to reliability indices, because the variance components are easier to interpret, with understandable units.
Content may be subject to copyright.
K. Chia, M. Sangeux
Quantifying sources of variability in gait analysis
Ko hl et h Chia a , b, * , Morgan Sa ng eu x a , b, c
a Hugh Williamson Gait Analysis Laboratory, Royal Children’s Hospital, Melbourne, Australia
b Murdoch Childrens Research Institute, Melbourne, Australia
c Department of Mechanical Engineering, University of Melbourne, Australia
* Corresponding author. Email: kohleth.chia@mcri.edu.au, kohleth@gmail.com. Address: MCRI, 50
Flemington Road, Parkville VIC 3052, Australia.
Abstract
Measurements from gait analysis are affected by many sources of variability. Schwartz et al. [1]
illustrated an experimental design and methods to estimate these variance components. However,
the derivation contains errors which could severely bias the estimation of some components.
Therefore, in this paper, we presented correction to this method using ANOVA and Likelihood
methods. Furthermore, we demonstrated how commonly used reliability indices like CMC and ICC
may be derived from the variance components. We advocate the use of the variance components, in
preference to reliability indices, because the variance components are easier to interpret, with
understandable units.
Acknowledgement
We thank the two anonymous reviewers for their helpful feedback, which enriched the content of this
paper.
Quantifying sources of variability in gait analysis
DOI: 10.1016/j.gaitpost.2017.04.040 12 May. 17
K. Chia, M. Sangeux 1
Introduction
Measurements from gait analysis are variable. The sources of the variability may be intrinsic or
extrinsic. Intrinsic variability corresponds to the variability of the subject under investigation, for
example the variability between strides of the same individual, or between individuals [2]. Extrinsic
variability corresponds to the variability of the gait analysis measurement process, for example marker
replacement between sessions, and between assessors, or different marker placement protocols and
processing workflows. Having the ability to differentiate and quantify these different sources of
variability, or as we call it, variance components, is important for estimating the reliability and
repeatability of gait analysis, comparing different methods and protocols, training assessors, and
sharing data between laboratories. To this purpose, three statistics are commonly used in the
literature, namely, the Coefficient of Multiple Correlation (CMC) as defined in Kadaba et al. [3], the
Intra-Class Correlation coefficient (ICC) [4], and the explicit quantification of variances following the
method proposed by Schwartz et al. [1].
CMC has been a popular choice because it is designed to handle curves rather than point data.
However, several authors have highlighted issues with CMC, such as its strong dependence on sample
size, or range of motion (ROM) [2, 5]. As we will elaborate below, we concur with Røislien et al. [5]
that CMC in its current form should not be used in these studies. ICC is a similar index but works on
the individual time point. However, as we will show below, both these indices may be derived from
the variance component estimates themselves. Therefore, in our view, the most appropriate and
fundamental framework is that of Schwartz et al. [1], which estimate the variance components
directly.
The method of Schwartz et al. [1] has been adopted in several studies [612]. However, their proposed
variance component estimators are biased, and as we will demonstrate, this bias may be severe. Our
primary objective is to present a corrected set of variance component estimators. In addition, we will
present methods to derive the ICC and CMC from the estimated variance components. Finally, we will
discuss methods to move beyond point-based calculation for curve data.
Materials & Methods
1. Data
The data we used to illustrate our methods was presented in Schwartz et al. [1]. While we did not have
access to the original data, figure 4 in Schwartz et al. [1] presents the inter-Trial, inter-Session, and
inter-Therapist standard deviation for 11 joint angles. We extracted the information from these curves
using Engauge Digitizer [13], which exported the coordinates of an irregular set of points on each
curve. These points were subsequently fitted with a natural B-spline [14]. The fitted splines were
plotted and visually inspected for significant discrepancy with the published curves. Finally, we
resampled 101 points from these splines at time points t=0, …, 100 to form our dataset.
2. Experimental Design
In general, to estimate variance components, one needs to conduct experiments that have multiple
realizations of each factor (therapist, session, trial etc.). These factors may have two kinds of
relationship with each other, crossed or nested, and it is important to differentiate between them.
Quantifying sources of variability in gait analysis
DOI: 10.1016/j.gaitpost.2017.04.040 12 May. 17
K. Chia, M. Sangeux 2
Factors A and B are crossed, if all possible combination between them occurs in the experiment, and
their effect remains the same regardless of the other variable. For example, in the design of Schwartz
et al. [1], the factors subject and therapist are crossed, because all subject-therapist combination
existed in the experiment, andthe effect of subject i is always the same regardless of which therapist
was doing the assessment. In addition, when two factors are crossed, the possibility for interaction
between factors arises. In contrast, when factors A and B are nested, there is a natural hierarchical
relationship between them, and the effect of the inner factor will depend on the level of the outer
factor. For example, the factor trial is nested within the factor session,because the effect of trial l in
session k will not be the same as that in session k', where k≠k'.
We need to differentiate between crossed and nested factors in order to account for all possible
sources of variability, including that from the interaction effect. Therefore, the accurate description
of the experimental design of Schwartz et al. [1] would be that 2 subjects are crossed with 4 therapists,
and within each subject-therapist combination, there are 3 sessions, and 5 trials nested within each
session.
Similarly, when representing the experimental design by figures, for example like in [1,6,7], it may be
informative to differentiate graphically between crossed and nested factors. The experimental design
of Schwartz et al. [1] (figure 2 in the original paper) may be redrawn to highlight the structure of the
factors in figure 1.
Figure 1:Graphical representation of the experimental design.
In gait analysis experiments, the degrees of freedom for the innermost factor (i.e. trial) is high, while
that for the outermost factors (subject or therapist) is low. For example, in our dataset, there are 4
therapists, thus 3 degrees of freedom for the estimation of the inter-Therapist variability. This is akin
Quantifying sources of variability in gait analysis
DOI: 10.1016/j.gaitpost.2017.04.040 12 May. 17
K. Chia, M. Sangeux 3
to estimating a variance from 4 numbers, which is inadequate. Therefore, it is often better to increase
the number of subjects and therapists (or just therapists if inter-Subject variability is not of interest),
while decreasing the number of trials to offset the increase in the resources demanded. Of course, in
practice, it is less costly to increase the number of trials than the number of therapists, so a trade-off
between cost and precision needs to be made.
3. Variance Estimation
Below, we present the original biased estimators of variance by Schwartz et al. [1], the corrected
version, as well as the ANOVA method of estimating variance components.
3.1. Schwartz et al. estimators of variance
Let  denotes the gait measurement (scalar) at time point , for subject , assessed
by therapist , at session  in trial . For notational simplicity we will
drop the  notation, except in cases where it might cause confusion.
The various means are defined as followed:
Session mean:


Subject-Therapist (Interaction) mean:


Therapist mean:

Subject mean:

Overall mean:

Schwartz et al. [1] defined the estimator for the variance components as followed:
Inter-trial: 



Inter-session:



Inter-therapist: 


However, these are biased for the target quantity 
, where  can be , , or .
Table 1 shows the expected value of these estimators, assuming the linear model as described in
section 2.3.3. The quantity 
denotes the variance components associated with the
interaction effect between a subject and a therapist. Unless we have strong reason to believe there is
no such effect, it is prudent to include it in the model.
Quantifying sources of variability in gait analysis
DOI: 10.1016/j.gaitpost.2017.04.040 12 May. 17
K. Chia, M. Sangeux 4
Table 1 Schwartz et al. estimators and their expected value.
Source
Schwartz’s estimators
Expected value of Schwartz’s estimators
Therapist
(target 






 












Session
(target 











Trial
(target 






There were two sources of bias in the estimators. Firstly, it over-estimated the degrees of freedom
associated with the target quantity. The degrees of freedom was always set to ,
whereas the true degrees of freedom was less for all factors. This reduction of degrees of freedom
reflects the fact that we do not know the various true means used in the estimators, but have to
estimate them from the data. For example, in the estimation of the inter-Trial variability (bottom row
in table 1), while we have used  many data, , in the sums of squares, we also estimated 
many
, therefore the correct degrees of freedom was .
Secondly, the expectation of the estimators was a linear combination of the variance components
rather than the target variance component itself. For example, the expected value of the Schwartz et
al. estimator for the inter-Session variability is not 
, but a linear combination of that and 
.
The reason for the appearance of the latter variance component is that the mean involved in the
estimators,
, is itself estimated by averaging the data at the innermost level, , and therefore
is associated with variability. In summary, these estimators will slightly under-estimate the innermost
variance components (inter-Trials), but over-estimate the rest, with increasing severity.
Table 1 provides a means to correct these estimators retrospectively for published literature.
Assuming there was no interaction effect (i.e. it was not estimable with the given data), Table 1
provides a system of 3 linear equations to calculate the unbiased estimates of the true variance
components: 
, 
, 
.
3.2. ANOVA
The relevant ANOVA method we employed is the random effect ANOVA, where we assume the factors
are a random quantity associated with the variance parameter 
. This method is in spirit similar
to that of Schwartz et al., except that it uses a different set of equations. A full ANOVA table associated
with the experimental design presented in figure 1 is shown in table 2.
Quantifying sources of variability in gait analysis
DOI: 10.1016/j.gaitpost.2017.04.040 12 May. 17
K. Chia, M. Sangeux 5
Table 2 ANOVA table for the experiment
Sum of Squares ()
Degree of
freedom ()
Expected Mean Square
()
































If we believe there was no interaction effect, the Interaction and Session strata may be combined, and
the combined values are:


 , , and 

.
By equating the mean squares () with their expectation (EMS), we obtain a system of
linear equations, which solves to the required variance components.
3.3. Maximum Likelihood
The ANOVA estimators are method-of-moments estimators. One drawback of this method is that it
requires the dataset to be balanced. Fortunately, both theoretical and computational advances have
since led to an alternative way of estimating variance components which does not require a balanced
dataset. This method is maximum likelihood estimation.
Implicitly assumed in the ANOVA method is the following linear mixed model:
   (1)
where denotes the overall mean, and the remaining terms denote the respective random effects,
each following an independent normal distribution with zero mean and variance 
for their
respective . Instead of computing the sums of squares, maximum likelihood method estimates
all unknown parameters directly by maximizing the likelihood function implied by the linear mixed
model. Due to this different approach of estimation, the estimated values may be slightly different to
Quantifying sources of variability in gait analysis
DOI: 10.1016/j.gaitpost.2017.04.040 12 May. 17
K. Chia, M. Sangeux 6
that from the ANOVA method. The actual calculation involved is beyond the scope of this paper, but
we point interested readers to Brown & Prescott [15]. The associated standard errors are also available
through the matrix of second derivatives, but it should be warned that they provide limited value due
to the asymmetry of the estimated variance components. Better alternatives for estimating
uncertainties exist but they require more involved computation. We refer interested readers to [15].
In practice, a variant of the maximum likelihood method, called residual maximum likelihood (REML)
[16] is preferred due to its ability to correct for the bias in the traditional maximum likelihood method.
3.4. Clinical usability of variance components
The estimated variance components can be used quite flexibly. For example, as we will show in section
2.4, they can be used to derive reliability indices like CMC and ICC. Alternatively, following Schwartz
et al.[1], we can compute the ratio 

. This ratio is useful because 
represents the
intrinsic variability that we cannot control, and thus serves as a good benchmark figure.
For comparison with the ratios computed in section 3 of Schwartz et al. [1], we have computed the
ratio of the extrinsic to intrinsic variation defined as:



The principal advantage of the variance components is that they can be used to derive quantities for
situations which have a completely different setup to the ones they were derived from. We give a
scenario to illustrate the point below.
Suppose a subject had two gait analysis visits, one use of the variance components is to estimate the
variability of the change between the two sessions. Using the estimated variance components, we can
calculate that, if the therapist was the same, then the variance of change would be:
 


whereas if the therapists were different it becomes:
 



So using different therapists increases variability and thus reduces our clinical precision.
Another use is to do sample size calculation. That is, calculate the required in order to achieve a
required threshold of variability, , by solving the following equation:
 


which leads to


Quantifying sources of variability in gait analysis
DOI: 10.1016/j.gaitpost.2017.04.040 12 May. 17
K. Chia, M. Sangeux 7
Yet another use is to simply inspect which of 

, and 
 is the largest, and then
target our managerial effort to reduce that source of variability.
From this scenario, it is worth noticing that, the setup of repeated visits, with or without the same
therapist, is different to the one from which the variance components were estimated. So once we
have estimated the variance components, they can be used in many different contexts.
4. Reliability indices
The variance components are the most fundamental quantities and below we illustrate how ICC and
CMC can be derived from them.
4.1. Intra-Class Correlation (ICC)
Recall that implicit in our estimation is the linear mixed model (1). The ICC is defined as the correlation
between two measurements that have exactly the same components, except for the ‘class’ variable.
For example, the ICC for therapist, is the correlation between two measurements from the same
subject, session, and trial, but different therapist. This quantity turns out to be a simple proportion of
the relevant variance components (proof in appendix):







Using our dataset, we have computed , but since 
is unknown, we removed this
term from both the numerator and denominator of the formula. Therefore, the  we
computed is 




.
The above quantity is useful for, for example, the training of a team of therapists, where the objective
is to have an ICC as close to 1 as possible. In practice, measurements from different therapists also
implies they are obtained from different sessions and trials. So, the ICC for inter-Therapist-Session-
Trial might also be of interest to, say, data quality control, and is defined as





4.2. Coefficient of Multiple Correlation (CMC)
Kadaba et al. [3] originally defined CMC in terms of various mean squares. In his paper, the
experiments involve multiple days and runs, instead of sessions and trials. Therefore, we have used a
different set of terms in this section. But statistically speaking, their ‘days’ is equivalent to our
‘sessions’ and their ‘runs’ is equivalent to our ‘trials’. So, consider an experiment where a gait variable
was measured on runs nested within days at time points . Kadaba et al. defined the intra-
day CMC to be:







and the inter-day CMC to be:







The idea behind such definitions is that the two numerators should capture the variance between the
runs and the days respectively. An argument similar to section 2.3.1 should indicate that the
Quantifying sources of variability in gait analysis
DOI: 10.1016/j.gaitpost.2017.04.040 12 May. 17
K. Chia, M. Sangeux 8
numerator for the inter-day CMC is a biased estimate of its target quantity. However, if we assume
the following linear model 
where is the overall mean, and the rest are random effects for the time point, day, and run
respectively. Then it is possible to derive a CMC based on the variance components:
















The reason why CMC is dominated by the range of movement (ROM) [2,5] is because CMC considers
the inter-time point difference as a variability,, which we believe is inappropriate. As a result, for
joints with large ROM, will be a lot larger than all the other variance components. Nonetheless, we
have shown how the CMC can be derived from the estimated variance components.
Results
Figure 2 plots the variance component curves as found in figure 4 of Schwartz et al. [1], together with
the corrected estimates (same as the ANOVA estimates) using methods outlined above. In addition,
the square root of the mean of the ANOVA curve, representing the average standard deviation, is also
reported. Finally, the ratio of extrinsic to intrinsic variation (see 2.3.4) is reported in the left panel. This
ratio will be slightly different to the ratio that would have been calculated using the individual average
standard deviations. This is due to the difference in the order of the mathematical operations involved
in the calculation (average then square-root vs square root then average). From the figure, we can see
that the bias of Schwartz et al. method [1] increases from inter-Trial, inter-Session, to inter-Therapist.
In particular, the original estimators usually over-estimate the inter-Therapist variability, and in some
cases, quite severely so. In cases such as Knee Rotation and Ankle Dorsiflexion, the corrected (ANOVA)
estimates show that the inter-Therapist variability is no longer the dominant source of variability.
Quantifying sources of variability in gait analysis
DOI: 10.1016/j.gaitpost.2017.04.040 12 May. 17
K. Chia, M. Sangeux 9
Figure 2 Estimated variance component in standard deviation (degree) using both Schwartz et al. original estimators and
the ANOVA estimators. Both time points estimate (curves) and average over the gait cycle (ANOVA only, top left hand
corner) are provided. The ratio of extrinsic to intrinsic variability, r (cf. section 2.3.4), is provided for each angle.
Quantifying sources of variability in gait analysis
DOI: 10.1016/j.gaitpost.2017.04.040 12 May. 17
K. Chia, M. Sangeux 10
Figure 3 Estimated Intra Class Correlation (ICC) for therapist using both Schwartz et al. estimator and the ANOVA estimator.
Quantifying sources of variability in gait analysis
DOI: 10.1016/j.gaitpost.2017.04.040 12 May. 17
K. Chia, M. Sangeux 11
Discussion
We have presented ANOVA based, and maximum likelihood based, methods to estimate and correct
the variance components as produced by Schwartz et al. [1]. In addition, we have presented methods
to derive reliability indices such as ICC and CMC from the estimated variance components. We have
also shown that the original uncorrected estimator severely over-estimated the variance components
at the higher level, which led to a more severely under-estimated ICC.
Interestingly, the ANOVA style of estimating variance components was found in McDowell et al [17],
and the maximum likelihood type method in McGinley et al [18,19]. However, these methods are not
widely adopted by the gait analysis community. Instead, it is quite common to see CMC [3] being used
as reliability measure. We believe the variance components should always be the fundamental
quantities to be computed when investigating issues related to variability, reliability, or repeatability.
This is because the variance components are actually variances which are well-understood statistical
quantities in meaningful units (degree2). In addition, reliability indices such as ICC and CMC can be
derived from these variances. Furthermore, they are applicable to a wide range of scenarios.
So far, we have shown how to estimate the variance components at individual time points. However,
the interest may be to obtain a variability measure for the entire curve as one number. Consider the
curve as represented by the vector , the variance of is naturally a
covariance matrix , where the  entry is the covariance . Under the
linear model (1), without interaction, can be decomposed into  . We
then seek a single number summary for each of the . Borrowing from the literature on optimal
experimental design [20], some of the common options are the trace and the determinant of the
covariance matrix. The determinant requires the entire covariance matrix to be estimated, which will
be difficult in practice. The trace, however, sums the variance at the individual time points (diagonal
terms), and is therefore equivalent to the average variance [2]. Therefore, purely for practicality we
recommend averaging the variance components at each time point to arrive at a curve based measure.
Quantifying sources of variability in gait analysis
DOI: 10.1016/j.gaitpost.2017.04.040 12 May. 17
K. Chia, M. Sangeux 12
References
[1] M.H. Schwartz, J.P. Trost, R.A. Wervey, Measurement and management of errors in
quantitative gait data, Gait Posture. 20 (2004) 196203.
[2] M. Sangeux, E. Passmore, H.K. Graham, O. Tirosh, The gait standard deviation, a single
measure of kinematic variability, Gait Posture. 46 (2016) 194200.
[3] M.P. Kadaba, H.K. Ramakrishnan, M.E. Wooten, J. Gainey, G. Gorton, G.V.B. Cochran,
Repeatability of Kinematic, Kinetic, and EMG Data in Normal Adult Gait.pdf, J. Orthop. Res. 7
(1989) 849860.
[4] J.L. Fleiss, The Design and Analysis of Clinical Experiments, John Wiley & Sons, Inc., Hoboken,
NJ, USA, 1999.
[5] J. Røislien, Skare, A. Opheim, L. Rennie, Evaluating the properties of the coefficient of multiple
correlation (CMC) for kinematic gait data, J. Biomech. 45 (2012) 20142018.
[6] K. Deschamps, F. Staes, H. Bruyninckx, E. Busschots, E. Jaspers, A. Atre, K. Desloovere,
Repeatability in the assessment of multi-segment foot kinematics, Gait Posture. 35 (2012)
255260.
[7] P. Salvia, S.V.S. Jan, A. Crouan, L. Vanderkerken, F. Moiseev, V. Sholukha, C. Mahieu, O.
Snoeck, M. Rooze, Precision of shoulder anatomical landmark calibration by two approaches:
A CAST-like protocol and a new anatomical palpator method, Gait Posture. 29 (2009) 587
591.
[8] K. Deschamps, F. Staes, H. Bruyninckx, E. Busschots, G.A. Matricali, P. Spaepen, C. Meyer, K.
Desloovere, Repeatability of a 3D multi-segment foot model protocol in presence of foot
deformities, Gait Posture. 36 (2012) 635638.
[9] K. Deschamps, P. Roosen, H. Bruyninckx, K. Desloovere, P.A. Deleu, G.A. Matricali, L. Peeraer,
F. Staes, Pattern description and reliability parameters of six force-time related indices
measured with plantar pressure measurements, Gait Posture. 38 (2013) 824829.
[10] P. Caravaggi, M.G. Benedetti, L. Berti, A. Leardini, Repeatability of a multi-segment foot
protocol in adult subjects, Gait Posture. 33 (2011) 133135.
[11] M. Manca, A. Leardini, S. Cavazza, G. Ferraresi, P. Marchi, E. Zanaga, M.G. Benedetti,
Repeatability of a new protocol for gait analysis in adult subjects, Gait Posture. 32 (2010) 282
284.
[12] K. Kaufman, E. Miller, T. Kingsbury, E. Russell Esposito, E. Wolf, J. Wilken, M. Wyatt, Reliability
of 3D gait data across multiple laboratories, Gait Posture. 49 (2016) 375381.
[13] M. Mitchell, Engauge Digitizer, (2014).
[14] C. De Boor, A practical guide to splines, Springer-Verlag New York, 1978.
[15] H. Brown, R. Prescott, Applied Mixed Models in Medicine, John Wiley & Sons, Ltd, Chichester,
UK, 2014.
[16] H.D. Patterson, R. Thompson, Recovery of inter-block information when block sizes are
unequal, Biometrika. 58 (1971) 545554.
[17] B.C. McDowell, R. Baker, V. Hewitt, a. Nurse, T. Weston, T. Dusoir, The variability of
goniometric measurements in children with spastic cerebral palsy, Gait Posture. 10 (1999) 64.
[18] J. McGinley, R. Baker, R. Wolfe, Quantification of kinematic measurement variability in gait
analysis, Gait Posture. 24 (2006) 5556.
[19] J. McGinley, R. Wolfe, M. Morris, M.G. Pandy, R. Baker, Variability of walking in able-bodied
adults across different time intervals, J. Phys. Med. Rehabil. Sci. 17 (2014) 610.
[20] A. Atkinson, A. Donev, R. Tobias, Optimum Experimental Designs, With SAS, Oxford university
press, Oxford, 2007.
Quantifying sources of variability in gait analysis
DOI: 10.1016/j.gaitpost.2017.04.040 12 May. 17
K. Chia, M. Sangeux 13
Appendix
Proof that ICC is a function of the estimated variance components
We prove the following equation only, all other ICCs equalities can be proved in a similar manner.






First, recall the linear model without interaction,
 
Then,  
So the variance is, 


(2)
But we also know that,

Therefore,




 (3)
Equating (3) and (2) we get 
And dividing both sides by the total variance 



concludes our proof.
... The mean value represents the average recognition rate in the gait phase, and the standard error describes the average difference between the mean recognition rate of different subjects and the mean overall recognition rate. Variance is also an indicator of statistics, which represents the deviation degree between sample and the mean [60]. Finally, the average recognition rate of the five gait results is calculated on the basis of the arithmetic mean. ...
... In this section, we calculate the variance of the recognition rate including DBN and SSA-DBN. The variance represents the deviation degree between the sample recognition rate and the average recognition rate, that is, the variance represents the stability of recognition [60]. ...
Article
Full-text available
In this paper, a gait classification method based on the deep belief network (DBN) optimized by the sparrow search algorithm (SSA) is proposed. The multiple features obtained based on surface electromyography (sEMG) are fused. These functions are used to train the model. First, the sample features, such as the time domain and frequency domain features of the denoised sEMG are extracted and then the fused features are obtained by feature combination. Second, the SSA is utilized to optimize the architecture of DBN and its weight parameters. Finally, the optimized DBN classifier is trained and used for gait recognition. The classification results are obtained by varying different factors and the recognition rate is compared with the previous classification algorithms. The results show that the recognition rate of SSA-DBN is higher than other classifiers, and the recognition accuracy is improved by about 2% compared with the unoptimized DBN. This indicates that for the application in gait recognition, SSA can optimize the network performance of DBN, thus improving the classification accuracy.
... Following the methodology proposed by Chia and Sangeux (2017), ICC estimates were then obtained as follows: ...
Article
Imaging techniques in anatomy have developed rapidly over the last decades through the emergence of various 3D scanning systems. Depending on the dissection level, non‐contact or tactile contact methods can be applied on the targeted structure. The aim of this study was to assess the inter and intra‐observer reproducibility of an ArUco‐based localisation stylus, that is, a manual technique on a hand‐held stylus. Ten fresh–frozen, unembalmed adult arms were used to digitalise the glenoid cartilage related to the glenohumeral joint and the contour of the clavicle cartilage related to the acromioclavicular joint. Three operators performed consecutive digitalisations of each cartilage contour using an ArUco‐based localisation stylus recorded by a single monocular camera. The shape of each cartilage was defined by nine shape parameters. Intra‐observer repeatability and inter‐observer reproducibility were computed using an intra‐class correlation (ICC) for each of these parameters. Overall, 35.2 ± 2.4 s and 26.6 ± 10.2 s were required by each examiner to digitalise the contour of a glenoid and acromioclavicular cartilage, respectively. For most parameters, good‐to‐excellent agreements were observed concerning intra‐observer (ICC ranging between 0.81 and 1.00) and inter‐observer (ICC ranging between 0.75 and 0.99) reproducibility. To conclude, through a fast and versatile process, the use of an ArUco‐based localisation stylus can be a reliable low‐cost alternative to conventional imaging methods to digitalise shoulder cartilage contours.
... Also, due to the EMG normalization, intermuscle amplitude comparisons were not considered by the SVM analysis, which could have limited its performance [3]. Second, we could not exclude that the observed signatures were partly due to extrinsic variability [32], i.e., variability in sensor placement or skin preparation. However, the same experienced operator managed all the measurements following identical guidelines between trials [15]. ...
... Reliability studies have quantified the extrinsic variability (Chia and Sangeux, 2017;McGinley et al., 2009;Schwartz et al., 2004), i.e., intersession and inter-examiner, but did not address the accuracy of the gait model against a benchmark. Investigations on accuracy focused on the constitutive elements of the gait model. ...
Article
There is a lack of knowledge about the accuracy of the Conventional Gait Model (CGM), compared to the true bone motion. Accuracy is hindered by both marker misplacement and soft-tissue artefact (STA). The effect of the lateral knee marker (KNE) misplacement and STA was determined from a secondary analysis of 13 subjects equipped with a total knee prothesis for which simultaneous dual-plane fluoroscopy and marker-based motion capture was available. In average, STA alone led to 3.3°, 2.9° and 6.7° errors for knee flexion, knee abduction, and the absolute hip rotation respectively. In comparison, marker misplacement led to 0.9°, 4.0° and 12.3° errors for the same kinematics. We showed that STA alone may lead to knee flexion-adduction cross-talk. This finding has clinical repercussions for the use of knee cross talk as a qualitative indicator of knee axis alignment. Our study showed that cumulative effects of marker misplacement and STA affect the transverse plane angles, making challenging to track internal/external rotation with less than 5° of errors.
... The Pearson correlation coefficients were used to estimate the correlation between marker placement precision or segment orientation with kinematic variability with p-value< 0.05 defined as statistically significant. Finally, one-way ANOVA was used to compare the intra-evaluator variability among the different levels of experience of the evaluators, with statistical significance set at p < 0.05 [21]. The experience was evaluated as four ordinal variables. ...
Article
Full-text available
Background: Gait analysis relies on the accurate and precise identification of anatomical landmarks to provide reliable and reproducible data. More specifically, the precision of marker placement among repeated measurements is responsible for increased variability in the output gait data. Research question: The objective of this study was to quantify the precision of marker placement on the lower limbs by a test-retest procedure and to investigate its propagation to kinematic data. Methods: The protocol was tested on a cohort of eight asymptomatic adults involving four evaluators, with different levels of experience. Each evaluator performed, three repeated marker placements for each participant. The standard deviation was used to calculate the precision of the marker placement, the precision of the orientation of the anatomical (segment) coordinate systems, and the precision of the lower limb kinematics. In addition, one-way ANOVA was used to compare the intra-evaluator marker placement precision and kinematic precisions among the different levels of the evaluator's experience. Finally, a Pearson correlation between marker placement precision and kinematic precision was analyzed. Results: Results have shown a precision of skin markers within 10 mm and 12 mm for intra-evaluator and inter-evaluator, respectively. Analysis of kinematic data showed good to moderate reliability for all parameters apart from hip and knee rotation that demonstrated poor intra- and inter-evaluator precision. Inter-trial variability was observed reduced than intra- and inter-evaluator variability. Moreover, experience had a positive impact on kinematic reliability since evaluators with higher experience showed a statistically significant increase in precision for most kinematic parameters. However, no correlation was observed between marker placement precision and kinematic precision which indicates that an error in the placement of one specific marker can be compensated or enhanced, in a non-linear way, by an error in the placement of other markers.
... According to the COSMIN checklist 1 , reliability was assessed using ICC and measurement error with SDC. The ICC is defined as the correlations between two measurements that have exactly the same components except for the 'class' component 15 . Variance of class components was computed from a single measure, two-way mixed effects model 16,17 with the lme4 package 18 of R 19 . ...
Article
Full-text available
Bi-plane X-ray provides 3D measurements of the lower limb based on the identification of anatomical landmarks in sagittal and frontal X-rays. In clinical practice, such measurements involve multiple operators and sessions. This study aimed at evaluating the reliability of anatomical landmarks identification and geometric parameters of the pelvis and femur measured with bi-plane X-rays before and after total hip arthroplasty (THA). Twenty-eight patients undergoing primary THA were selected retrospectively. Two operators performed three reconstructions for each patient before and after THA. Intraclass correlation (ICC) and smallest detectable change (SDC) were computed for intra-operator, inter-operator, and test–retest conditions. Most anatomical landmark positions had good to excellent SDC (< 5 mm) apart from the centre of the sacral slope, greater trochanter, and anterior superior iliac spines (up to 7.1, 16.9, and 21.5 mm respectively). Geometric parameters had moderate to excellent SDC, apart from femoral and stem torsion, pelvic incidence, and APP inclination with poor SDC (9–12°). The sagittal view had significantly higher measurement errors than the frontal view. Test–retest and inter-operator conditions had no significant differences suggesting a low influence of patient posture. Osteoarthritis and the presence of implants did not seem to influence reliability and measurement error. This study could be used as a reference when assessing lower limb structure with bi-plane X-rays.
... The inter-session and inter-operator repeatability was computed on the angle error , following the recommendations of Schwartz et al. [20] and the corrections of Chia and Sangeux [21]. The exact computation is provided in Appendix B. ...
Article
A pre-requisite for obtaining human movement kinematics from Inertial Measurement Units (IMUs) is to define the relative orientation between the IMU coordinate systems and that of the underlying segment: a step called sensor-to-segment (S2S) calibration. Many S2S calibration methods have been proposed in the literature for the lower-body segments. However, these methods were not compared in a single study and the methodology differences between the studies make an objective assessment of the proposed S2S methods impossible. The present study aims firstly to compare different S2S calibration methods of the lower limbs in terms of accuracy, repeatability, and usability, and secondly to specify the impact of the S2S calibration method on lower-body kinematics. In Experiment 1, two postures, eight active, and five passive movements, and the use of a device were tested to define lower-body segment axes. To isolate the effect of the S2S calibration methods from IMU measures, the IMU measure was mimicked using marker trajectories tracked by an optoelectronic system. The angles between the segment axes obtained with the S2S calibration methods and axes of reference based on an optoelectronic methodology were then compared during a study involving fifteen subjects. The results do not reveal unique methods, but enable some to be discarded in terms of angle error and repeatability. Following these results, scoring of the methods was proposed in order to help select the most suitable S2S calibration methods given experimental context or subject ability. In Experiment 2, three kinematics obtained with IMUs after S2S calibrations combining the best previous methods, and kinematics obtained with the reference, were compared through RMSE, correlation coefficients, and difference in range of motion during gait, running, and cycling on an ergocycle. The combination, which included only, standing and sitting postures showed significantly lower RMSE (p<0.01) and range of motion differences (p<0.01) in ankle abduction/adduction and plantar/dorsiflexion respectively, making this combination a good candidate for S2S calibration of the lower-body for non-pathological subjects.
... The kinematic data are calculated from marker trajectories on a frame by frame basis 5 . Any measurement errors in gait analysis introduce variability into the output data and negatively impact data interpretation 6 . Marker placement has been reported as the primary cause of variability in gait analysis 7,8 . ...
Article
Full-text available
Clinical gait analysis supports treatment decisions for patients with motor disorders. Measurement reproducibility is affected by extrinsic errors such as marker misplacement?considered the main factor in gait analysis variability. However, how marker placement affects output kinematics is not completely understood. The present study aimed to evaluate the Conventional Gait Model's sensitivity to marker placement. Using a dataset of kinematics for 20 children, eight lower-limb markers were virtually displaced by 10mm in all four planes, and all the displacement combinations were recalculated. Root-mean-square deviation angles were calculated for each simulation with respect to the original kinematics. The marker movements with the greatest impact were for the femoral and tibial wands together with the lateral femoral epicondyle marker when displaced in the anterior-posterior axis. When displaced alone, the femoral wand was responsible for a deviation of 7.3° (±?1.8°) in hip rotation. Transversal plane measurements were affected most, with around 40% of simulations resulting in an effect greater than the acceptable limit of 5°. This study also provided insight into which markers need to be placed very carefully to obtain more reliable gait data.
Article
Full-text available
The identification of relevant and valid biomarkers to distinguish patients with non-specific chronic low back pain (NSCLBP) from an asymptomatic population in terms of musculoskeletal factors could contribute to patient follow-up and to evaluate therapeutic strategies. Several parameters related to movement impairments have been proposed in the literature in that respect. However, most of them were assessed in only one study, and only 8% were evaluated in terms of reliability, validity and interpretability. The aim of this study was to consolidate the current knowledge about movement biomarkers to discriminate NSCLBP patients from an asymptomatic population. For that, an experimental protocol was established to assess the reliability, validity and interpretability of a set of 72 movement biomarkers on 30 asymptomatic participants and 30 NSCLBP patients. Correlations between the biomarkers and common patient reported outcome measures were also analysed. Four biomarkers reached at least a good level in reliability (ICC ≥ 0.75) and validity (significant difference between asymptomatic participants and NSCLBP patients, p ≤ 0.01) domains and could thus be possibly considered as valuable biomarkers: maximal lumbar sagittal angle, lumbar sagittal angle range of motion, mean lumbar sagittal angular velocity, and maximal upper lumbar sagittal angle during trunk sagittal bending. These four biomarkers demonstrated typically larger values in asymptomatic participants than in NSCLBP patients. They are in general weakly correlated with patient reported outcome measures, arguing for a potential interest in including related musculoskeletal factors in the establishment of a valuable diagnosis and in guiding treatment response.
Article
Background: Children with cerebral palsy develop foot deformities due to a combination of factors including muscle shortening, hypertonia, weakness, and cocontraction of muscles acting at the ankle joint resulting in an altered gait pattern. We hypothesized these factors affect the peroneus longus (PL) and tibialis anterior (TA) muscles couple in children who develop equinovalgus gait first followed by planovalgus foot deformities. Our aim was to evaluate the effects of abobotulinum toxin A injection to the PL muscle, in a cohort of children with unilateral spastic cerebral palsy and equinovalgus gait. Methods: This was a prospective cohort study. The children were examined within 12 months before and after injection to their PL muscle. Twenty-five children of mean age 3.4 (S.D.: 1.1) years were recruited. Results: We found significant improvement in foot radiology measures. Passive extensibility of the triceps surae did not change, whereas active dorsiflexion increased significantly. Nondimensional walking speed increased by 0.1 (95% confidence interval [CI], [0.07, 0.16]; P < 0.001), and the Edinburgh visual gait score improved by 2.8 (95% CI, [-4.06, -1.46]; P < 0.001). Electromyography showed increased recruitment for gastrocnemius medialis (GM) and TA but not for PL during the reference exercises (standing on tip toes for GM/PL, active dorsiflexion for TA) and decreased activation percentages for PL/GM and TA across sub-phases of gait. Conclusions: One key advantage of treating the PL muscle only might be to address foot deformities without interfering with the main plantar flexors that are instrumental to support body weight during gait.
Article
Full-text available
Measurement of gait kinematic variability provides relevant clinical information in certain conditions affecting the neuromotor control of movement. In this article, we present a measure of overall gait kinematic variability, GaitSD, based on combination of waveforms' standard deviation. The waveform standard deviation is the common numerator in established indices of variability such as Kadaba's coefficient of multiple correlation or Winter's waveform coefficient of variation.Gait data were collected on typically developing children aged 6-17 years. Large number of strides was captured for each child, average 45 (SD: 11) for kinematics and 19 (SD: 5) for kinetics. We used a bootstrap procedure to determine the precision of GaitSD as a function of the number of strides processed. We compared the within-subject, stride-to-stride, variability with the, between-subject, variability of the normative pattern. Finally, we investigated the correlation between age and gait kinematic, kinetic and spatio-temporal variability.In typically developing children, the relative precision of GaitSD was 10% as soon as 6 strides were captured. As a comparison, spatio-temporal parameters required 30 strides to reach the same relative precision. The ratio stride-to-stride divided by normative pattern variability was smaller in kinematic variables (the smallest for pelvic tilt, 28%) than in kinetic and spatio-temporal variables (the largest for normalised stride length, 95%). GaitSD had a strong, negative correlation with age. We show that gait consistency may stabilise only at, or after, skeletal maturity.
Article
Full-text available
When assessing reliability of three dimensional gait analysis, the coefficient of multiple correlation (CMC), a measure of similarity of waveforms, is frequently used. Several shortcomings of the method have been reported. It is strongly related to the range of motion (ROM); it is supposed to take values from 0 to 1, but is known to obtain complex values and break down. Removing offset before calculations appears to make it insensitive to apparent differences in gait. In the present work we use stochastic simulations to demonstrate why all of this happens, and the consequences thereof. We also demonstrate other, lesser-known, issues with the method, such as the strong dependency on number of subjects and test situations. The results are demonstrated on real data from an inter-rater repeatability study. We conclude that the CMC in its current form is generally unsuitable as a tool for assessing reliability in kinematic gait data.
Book
This book is based on the author's experience with calculations involving polynomial splines. It presents those parts of the theory which are especially useful in calculations and stresses the representation of splines as linear combinations of B-splines. After two chapters summarizing polynomial approximation, a rigorous discussion of elementary spline theory is given involving linear, cubic and parabolic splines. The computational handling of piecewise polynomial functions (of one variable) of arbitrary order is the subject of chapters VII and VIII, while chapters IX, X, and XI are devoted to B-splines. The distances from splines with fixed and with variable knots is discussed in chapter XII. The remaining five chapters concern specific approximation methods, interpolation, smoothing and least-squares approximation, the solution of an ordinary differential equation by collocation, curve fitting, and surface fitting. The present text version differs from the original in several respects. The book is now typeset (in plain TeX), the Fortran programs now make use of Fortran 77 features. The figures have been redrawn with the aid of Matlab, various errors have been corrected, and many more formal statements have been provided with proofs. Further, all formal statements and equations have been numbered by the same numbering system, to make it easier to find any particular item. A major change has occured in Chapters IX-XI where the B-spline theory is now developed directly from the recurrence relations without recourse to divided differences. This has brought in knot insertion as a powerful tool for providing simple proofs concerning the shape-preserving properties of the B-spline series.
Article
The aim of this study was to analyze the repeatability of gait analysis studies performed across multiple trials, sessions, and laboratories. Ten healthy participants (6 male/4 female, mean age of 30, mean BMI of 24 kg/m2) were assessed in 3 sessions conducted at each of the three Centers of Excellence for Amputee Care within the Department of Defense. For each test session, kinematic and kinetic parameters were collected during five walking trials for each limb. One independent examiner at each site placed markers on the subjects. Biomechanical data were collected at two walking speeds: self-selected and Froude speed. Variability of the gait data was attributed to inter-trial, inter-session, and inter-lab errors for each subject. These error sources were averaged across all ten subjects to obtain a pooled error estimate. The kinematic errors were fairly consistent at the two walking speeds tested. Median inter-lab kinematic errors were <5.0 degrees (median 2.3 degrees) for all joint angle measurements. However, the kinetic error differed significantly between walking speeds. The median inter-lab kinetic error for the self-selected speed was 0.112 Nm/kg (ICR 0.091-0.184) with a maximum of 0.226 Nm/kg. The errors were greatly reduced when the subjects walked at their Froude speed. The median inter-lab error was 0.048 Nm/kg (ICR 0.025–0.078, maximum 0.086). These data demonstrate that it is possible to get reliable data across multiple gait laboratories, particularly when gait speed is standardized across testing sessions. A key similarity between sites was the use of identical anatomical segment definitions for the respective gait models.
Article
Background: Functional interpretation of plantar pressure measurements is commonly done through the use of ratios and indices which are preceded by the strategic combination of a subsampling method and selection of physical quantities. However, errors which may arise throughout the determination of these temporal indices/ratio calculations (T-IRC) have not been quantified. The purpose of the current study was therefore to estimate the reliability of T-IRC following semi-automatic total mapping (SATM). Methods: Using a repeated-measures design, two experienced therapists performed three subsampling sessions on three left and right pedobarographic footprints of ten healthy participants. Following the subsampling, six T-IRC were calculated: Rearfoot-Forefoot_fti, Rearfoot-Midfoot_fti, Forefoot medial/lateral_fti, First ray_fti, Metatarsal 1-Metatarsal 5_fti, Foot medial-lateral_fti. Findings: Patterns of the T-IRC were found to be consistent and in good agreement with corresponding knowledge from the literature. The inter-session errors of both therapists were similar in pattern and magnitude. The lowest peak inter-therapist error was found in the First ray_fti (6.5 a.u.) whereas the highest peak inter-therapist error was observed in the Forefoot medial/lateral_fti (27.0 a.u.) The magnitude of the inter-session and inter-therapist error varied over time, precluding the calculation of a simple numerical value for the error. The difference between both error parameters of all T-IRC was negligible which underscores the repeatability of the SATM protocol. Conclusion: The current study reports consistent patterns for six T-IRC and similar inter-session and inter-therapist error. The proposed SATM protocol and the T-IRC may therefore serve as basis for functional interpretation of footprint data.
Article
Repeatability studies on 3D multi-segment foot models (3DMFMs) have mainly considered healthy participants which contrasts with the widespread application of these models to evaluate foot pathologies. The current study aimed at establishing the repeatability of the 3DMFM described by Leardini et al. in presence of foot deformities. Foot kinematics of eight adult participants were analyzed using a repeated-measures design including two therapists with different levels of experience. The inter-trial variability was higher compared to the kinematics of healthy subjects. Consideration of relative angles resulted in the lowest inter-session variability. The absolute 3D rotations between the Sha-Cal and Cal-Met seem to have the lowest variability in both therapists. A general trend towards higher σ(sess)/σ(trial) ratios was observed when the midfoot was involved. The current study indicates that not only relative 3D rotations and planar angles can be measured consistently in patients, also a number of absolute parameters can be consistently measured serving as basis for the decision making process.