You are on page 1of 14

Original Article

Proc IMechE Part P:


J Sports Engineering and Technology
1–14
Effect of mechanically aged minimalist Ó IMechE 2019
Article reuse guidelines:
and traditional footwear on female sagepub.com/journals-permissions
DOI: 10.1177/1754337118824001

running biomechanics journals.sagepub.com/home/pip

Nadine Lippa1,2, Jason Bonacci3, Paul K Collins4,


James W Rawlins2 and Trenton E Gould1

Abstract
The purpose of this study was to improve the understanding of human–material interactions by combining polymer engi-
neering and biomechanical approaches. The forefoot and heel of traditional shoes and minimalist running shoes were
degraded using a mechanical aging protocol to quantify (1) the effect of subject-specific degradation and (2) human biome-
chanical effects due to decreased material properties. Four recreational-level female participants ran in the shoes pre-
mechanical aging to determine the aging protocol input parameters and post-mechanical aging to evaluate the effect of
degradation on kinematics and kinetics. Initial biomechanics translated into different mechanical aging input parameters
among conditions: 500 greater number of impact cycles for minimalist shoe, 430 N higher peak force for forefoot, 75 kPa
greater peak stress for the heel, 3.1 and 13.7 kN/s greater loading rates for minimalist shoe and the heel, and recovery
time 220 ms greater for the heel. From mechanical aging, the shoe types and regions lost 1.2–1.8 mm thickness and 38%–
54% energy absorption overall, while drop decreased 0.6 mm for traditional shoe only. Samples degraded at different rates
depending on runner-specific input parameters. Human kinematics and kinetics were affected by both shoe type and aging.
Aging of the shoes decreased knee flexion velocity (1°/s; p = 0.01), decreased ankle dorsiflexion during stance (3°,
p = 0.01), and increased the vertical loading rate (4 BW/s, p = 0.01). The results support previous findings that different
footwear influence running biomechanics and concurrently advance footwear science to show running biomechanics are
also influenced by shoe degradation rates, such that unique and intuitive human–material interactions are apparent.

Keywords
Ethylene vinyl acetate foam, degradation, hysteresis, energy absorption, running shoes

Date received: 9 December 2017; accepted: 17 December 2018

Introduction rates, depending on the population studied.4–6,9–13


However, the knowledge of the midsole degradation
Significant engineering efforts have sought to advance rates in both traditional and minimalist shoes is sparse.
footwear technology through design or material devel- In general, mechanical aging (MA) of midsoles has
opment of the midsole.1–3 Running shoe midsoles are revealed that EVA foam thickness affects inherent
usually composed of ethylene vinyl acetate (EVA) foam
and function as the main energy damping component
of the shoe.1 While the majority of runners use tradi- 1
School of Kinesiology and Nutrition, The University of Southern
tional shoes (TS), the increase in popularity of barefoot Mississippi, Hattiesburg, MS, USA
2
running led to a rise in use of minimalist shoes (MS).4 School of Polymer Science and Engineering, The University of Southern
Mississippi, Hattiesburg, MS, USA
The main differences between traditional and minimal 3
School of Exercise & Nutrition Sciences, Deakin University, Waurn
shoes are the overall thickness of the midsole and the Ponds, Australia
‘‘drop,’’ or the difference in thickness between the heel 4
Mechanical Engineering, Deakin University, Waurn Ponds, Australia
and forefoot. Modifications of the midsole cushion pro-
file in traditional and minimalist shoes have been associ- Corresponding author:
Trenton E Gould, College of Education and Human Sciences, School of
ated with changes in running gait.4–8 For example, MS Kinesiology and Nutrition, The University of Southern Mississippi, 118
have been suggested to promote shorter strides, a more College Drive, #5023, Hattiesburg, MS 39406, USA.
distal footstrike, and either higher or lower loading Email: trent.gould@usm.edu
2 Proc IMechE Part P: J Sports Engineering and Technology 00(0)

energy absorption capability, with thick, compliant Methods


foam absorbing more energy overall, but degrading at a
faster rate than thinner, stiff foam.14 Additional work is
Participants
required to determine whether decreased drop, Four female recreational runners with a mean (6SD)
decreased energy absorption, and altered running gait age of 25.0 6 5.6 years, height of 1.69 6 0.07 m, and
occur in a more biofidelic degradation of the midsole body mass of 60.6 6 3.4 kg participated. The partici-
with realistic forces applied for each shoe-region condi- pants were heel strikers with US shoe sizes ranging
tion. It may be anticipated that important interactions from 7 to 8.5, running 21.0 6 4.2 km per week at a typi-
among shoe type, midsole degradation, and runner bio- cal pace of 5.7 6 0.4 min/km. Participants were
mechanics may occur.4–7 excluded if they had any illnesses that may affect bal-
Shoe degradation is known to occur15–22 and has ance, a lower limb injury in the past 6 months, or were
been associated with an increased risk for running pregnant.
injury,23 yet changes in traditional and minimalist shoe Participants were familiar with treadmill running
material properties with aging and its consequences on and were familiarized with the laboratory treadmill
human running biomechanics have not been reported. before participation.
One study reported that aged shoes affected running
kinematics (i.e. in worn shoes: less maximum lean and
less forward lean at toe-off for the torso; reduced maxi-
Experimental shoe conditions
mum dorsiflexion and increased plantarflexion at toe- Two pairs of New BalanceÒ running shoes were pur-
off),24 but the degree of shoe degradation was chased for each subject: (1) minimalist shoe model
unknown. Conversely, footwear material studies have WR10WW2 (MS) and (2) traditional shoe model
reported that shoe degradation decreases midsole thick- W880M13 (TS). To increase similarity between the two
ness, compliance, and the shoe’s ability to absorb shoe models (e.g. thickness, bending stiffness), the TS
energy,15–22 but have not examined how these changes shoes were slightly modified by (1) removing the
influence running biomechanics. For the running ath- optional foam insole and (2) removing the hard plastic
lete, a probable consequence of reduced thickness and Stability WebÒ. A 20-mm-diameter hole was cut in the
compliance is an increased loading rate,4,25 which has uppers of all shoes, directly above the metatarsal region
been reported as a risk factor for running overuse inju- of the footbed, to facilitate accurate measurement of
ries.26,27 However, there is no literature that has the forefoot region thickness with customized calipers
directly related thickness-dependent shoe energy throughout testing and to permit the compression pla-
absorption to runner kinematics, kinetics, or work ten access to the surface of the shoes during MA
done by the lower limb joints. (Figure 1).
The ability to endorse or refute the hypothesis that
new shoes ‘‘perform’’ better than aged shoes requires
bridging polymer science and biomechanics, which tend Footwear MA protocol
to occupy separate research silos. Consequently, previ- The experimental protocol consisted of a first round
ous footwear research falls into two categories: (1) human participant testing (Pre-MA), followed by MA,
human biomechanics assessment without reporting and then a second round of human participant testing
how properties of shoes changed over time4–7 or (2) (Post-MA) (Figure 2). The shoes were mechanically
shoe MA under nonspecific conditions of an ‘‘average’’ aged by a modified version of a previously developed
runner excluding the interrelationship of degradation protocol23,30–32 using a MTS 858 Servo Hydraulic
and human biomechanics.15–21,28,29 mechanical testing unit (Eden Prairie, MN; PDIF tun-
Our aim was to determine how footwear degradation ing factors: 36, 6.4, 0, and 0). Platens with 1 mm cham-
affects knee and ankle running biomechanics as a func- fer and 52 or 70 mm diameter applied forces to the heel
tion of footwear type. The following hypotheses were or forefoot shoe regions, respectively. Specific input
tested: (1) MA input parameters (defined based on parameters of the MA protocols were based on decon-
lower limb running kinematics and kinetics) will differ voluted bimodal waveforms corresponding to the four
between minimalist and traditional shoes, (2) the rate of participants’ Pre-MA kinetics such that the heelstrike
footwear degradation will be affected by footwear- peaks (F1) were applied to the heel regions and the
specific kinematics and kinetics, and (3) kinematics and active peaks (F2) were applied to the forefoot of each
kinetics will change as a function of polymer degrada- shoe. Therefore, four unique conditions were tested: TS
tion and will differ between traditional and minimalist heel (n = 8), TS forefoot (n = 8), MS heel (n = 8), and
shoes. The first hypothesis was substantiated by previ- MS forefoot (n = 8). Test duration and number of
ous work,4–7 but several gaps still exist. This study is cycles depended on the self-determined speed and gait
the first to take the multidisciplinary approach of con- of each subject in each shoe condition, such that MA
currently relating footwear degradation and human was personalized to each subject’s pre-aging gait. Stride
running biomechanics using established techniques and length (m/stride) was estimated by dividing the tread-
contemporary footwear conditions. mill speed (m/s) by the stride frequency (strides/s). The
Lippa et al. 3

Figure 1. Test setup for (a) an example MS sample in compression and (b) the 20-mm-diameter hole cut into an example TS upper.
The inferior portion of the platen is not visible because it was inside the shoe in direct contact with the footbed.

interaction with a change in footwear or surface


hardness.34
The participants ran on an instrumented treadmill
(Bertec, Columbus, OH) at a self-selected speed deter-
mined prior to testing. A digital video camera captured
the sagittal view during running trials. Kinematic and
kinetic data were collected for a 20-s duration during
the final 30 s of the 5-min run. The shoe conditions were
counterbalanced between participants and sessions to
eliminate ordering effects.

Kinematics and kinetics


Figure 2. Schematic representation of the study design. The Retroreflective markers were attached bilaterally across
triangle vertices represent data collection and analysis phases, 16 standard locations, fixed using double-sided adhe-
while the sides indicate how the outcome of each phase affects sive tape, and further secured with fixomull tape.35
the next. Pre-MA human testing informed input variables for Markers on the medial epicondyles of the knee and
mechanical aging. Over 42 km, MA caused footwear degradation,
medial malleolus were used only for calibration pur-
which affected Post-MA human variables.
poses and were removed prior to running trials. All
analyses were completed on the right limb. Marker tra-
jectories and ground reaction forces (GRF) were
stride length and force-time profile were then used to
recorded at a sampling frequency of 250 and 1500 Hz,
calculate the number of impacts needed to age the
respectively. A standard Vicon calibration procedure
shoes for 42 km. A marathon distance was chosen for
was followed.33,36–39
the study because measurable midsole degradation was
Marker trajectory and GRF data were low-pass fil-
shown to occur within this distance,14 and it is a stan-
tered (20 Hz) with a cubic smoothing spline within
dard distance for an athlete to run continuously.
Vicon Nexus, then exported to Visual 3D software (C-
Motion Rockville, MD). A kinematic model33,37 was
defined within Visual 3D from the participant’s stand-
Human testing protocol
ing calibration trial. Positive values represented posi-
Participants attended the gait laboratory on two occa- tions of knee flexion and ankle dorsiflexion. Negative
sions (Pre-MA and Post-MA) to run in each of the values represented positions of knee extension and
shoe conditions (e.g. MS-Pre represented running in ankle plantarflexion. Joint kinetics were calculated
the MS shoe on the first visit). Thirty-two retroreflec- from filtered data using a standard Newton–Euler
tive markers were attached to the pelvis and lower inverse-dynamics approach. Three-dimensional joint
limbs as per a previously validated musculoskeletal moments were expressed as internal moments normal-
model,33 and the same investigator performed the pla- ized to the product of body mass and height.40 Joint
cement of retroreflective markers on every participant. stiffness was calculated as the change in joint moment
Data collection consisted of two 5-min running trials divided by the change in joint angle from initial contact
wearing TS and MS footwear, with a 10-min rest to the point of greatest joint flexion.5 Power was the
between trials. Previous research has identified that product of the net joint moment and corresponding
4 min is sufficient time to optimize the foot–surface angular joint velocity. The negative work done during
4 Proc IMechE Part P: J Sports Engineering and Technology 00(0)

initial stance was computed for the knee and ankle shoe material variables, three 2 (shoe type: TS, MS) 3
joints by integrating the absorption portion of the 2 (region: forefoot, heel) 3 7 (distance: 0, 1, 3, 5, 10, 25,
power–time curve. Both power and work were normal- 42 km) tests were used. As a post hoc test, Bonferroni-
ized to body mass. corrected pairwise comparisons were implemented on
the within-subjects time variable to find specific differ-
ences in thickness and drop.
Data analysis Objective 3. To determine whether changes in kine-
The MA protocol input parameters calculated from matics and kinetics as a function of aging differed
each subject’s bimodal force–time data were F1 and F2 between traditional and minimalist shoes, eight 2 (TS,
peak forces (PF; N) and recovery time between impacts MS) 3 2 (Pre-MA, Post-MA) RM ANOVAs were cal-
(t; ms). From the input parameters and platen sizes, culated on the aforementioned human kinetics (four
loading rate (LR; BW/s) and peak stress (speak; kPa) variables) and kinematics (four variables).
were also calculated. Throughout aging, the main
dependent variables from the MA output were thick- For significant post hoc findings, effect size was cal-
ness, drop, net displacement (ND), energy absorbed culated using Cohen’s d. The a level was set a priori at
(EA), and percent energy absorbed (%EA). Drop was 0.05. For inferences about patterns of means across
the measured thickness difference between heel and conditions, within-subjects confidence intervals are gra-
forefoot. ND was the displacement maximum sub- phically presented on normalized scores for each sub-
tracted by the minimum for a single cycle. EA was the ject using the Masson and Loftus procedure.41 Results
integral of the force–displacement curve for a single are reported as mean (6 1 SD) unless otherwise noted.
cycle (equation (1)). Percent energy absorbed was calcu-
lated by equation (2), where m was a measurement time
point corresponding to 0, 1, 3, 5, 10, 25, and 42 km
Results
ð MA input parameters
Energy AbsorbedðEAÞ = Fdx ð1Þ All input variables are summarized in Table 1 and sig-
nificant statistics are summarized in Table 2. Each
marathon MA test lasted for 4.8 (0.1) h or 23,390 (406)
EAm
%EA = 3100% ð2Þ impact cycles. On average, MS underwent 500 more
EA0
impact cycles than TS. Peak forces were 430 N higher
for forefoot (F2 force) than heel (F1 force) (Figure 3).
Temporospatial stride characteristics, joint kine- Peak stress (speak) was 75 kPa greater for heel than fore-
matics, moments, power, and work were extracted for foot. The loading rates were 3.1 and 13.7 kN/s greater
statistical analysis using a customized MATLAB for MS than TS and heel than forefoot, respectively.
(MathWorks Inc., Natick, MA) program. The data for The shoe recovery times were 220 ms greater for heel
each participant were normalized to the stride cycle than forefoot. The resulting mechanical force–time pro-
(0%–100%), averaged over the 20 s for each trial and files closely resembled that of the participants’ impulse
plotted over the stance phase. Dependent variables of curves (Figure 3).
interest included the peak sagittal plane kinematics,
peak joint moments and power joint stiffness, and neg-
ative work done at the knee and ankle joints during the Footwear degradation
first phase of stance (absorption). Differences between All MA output variables are listed in Table 3 and sig-
experimental conditions were examined using factorial nificant statistics are summarized in Table 4. As a func-
repeated-measures analysis of variance (RM ANOVA). tion of human and MA testing, midsole thickness
decreased overall, but there were significant stages of
Objective 1. To determine whether minimalist and tra- degradation and recovery between trials (Figure 4(a)).
ditional shoes promoted unique MA input parameters Post hoc tests on thickness revealed that (1) all condi-
(i.e. kinematics and kinetics), the MA input variables tions decreased from Pre-MA to Post-MA and (2) all
were tested with five 2 (shoe type: TS, MS) 3 2 (region: conditions decreased from baseline to follow-up run-
forefoot, heel) RM ANOVAs. ning tests (Figure 4(a)). Comparing across shoe-
Objective 2. To determine whether the kinematics and thickness conditions, TS thickness was 8.4 mm greater
kinetics of each shoe type promoted unique footwear than MS and heel was 8.8 mm greater than forefoot.
degradation, five RM ANOVAs were used. Specifically, There was an interaction effect, which indicated that
for change in shoe thickness throughout testing, a 2 thickness decreased at a faster rate for heel (than fore-
(shoe type: TS, MS) 3 2 (region: forefoot, heel) 3 6 foot) and TS (than MS). As a function of aging dis-
(Time: Before/After baseline running test, Pre- and tance, drop decreased by 0–0.6 mm (Figure 4(b)).
Post-MA, Before/After follow-up running test) test was However, post hoc comparison tests on drop revealed
used. For change in shoe drop between TS and MS, a 2 that (1) MS drop did not significantly change from any
(shoe type: TS, MS) 3 6 (Time) test was used. For the treatment; (2) from Pre-MA to Post-MA, TS drop
Lippa et al. 5

Table 1. Group mean (SD) of MA input variables across shoe and region variables derived from MA input calculated from subject-
specific running trials.

MA input variables MS forefoot MS heel TS forefoot TS heel

Cycles [#] 23,629 (364) 23,629 (364) 23,125 (284) 23,125 (284)
Peak force [N] 1318 (94) 909 (127) 1297 (111) 851 (104)
Peak s [kPa] 342 (24) 428 (60) 337 (29) 401 (49)
Loading rate [N/s] 9048 (1007) 25,443 (3552) 8660 (1196) 19,608 (3124)
Recovery time [ms] 437 (4) 658 (18) 443 (8) 657 (16)

MA: mechanical aging; MS: minimalist shoe; TS: traditional shoe.

Table 2. Statistical results for MA input variables across shoe and region variables derived from MA input calculated from subject-
specific running trials.

MA input variables Result DoF F p f

Cycles (#) MS . TS 1, 28 23.10 \ 0.001 0.91


Loading rate (N/s) MS . TS 1, 28 12.48 0.001 0.67
Peak force (N) Forefoot . heel 1, 28 121.39 \ 0.001 0.91
Peak s (kPa) Heel . forefoot 1, 28 24.03 \ 0.001 2.09
Loading rate (N/s) Heel . forefoot 1, 28 240.97 \ 0.001 2.94
Recovery time (ms) Heel . forefoot 1, 28 3517.98 \ 0.001 11.14

MA: mechanical aging; MS: minimalist shoe; TS: traditional shoe; DoF: degree of freedom; f: the standardized mean difference expressed as Cohen’s f.

Figure 3. Representative Pre-MA subject plots for TS and MS show similar bimodal ground reaction force response (a). Mechanical
testing input for all shoe-region conditions demonstrated differences between each condition (b). TS heel and forefoot regions
plotted over 1.5 cycles demonstrate different loading rates, peak forces, and recovery times afforded to each shoe region (c).

decreased 0.6 mm; and (3) there were no differences in Comparing across shoe-thickness conditions, TS EA
drop measured from baseline to follow-up running tests was 233 mJ greater than MS, but there was no differ-
(Figure 4(b)). Comparing across shoe-thickness condi- ence between heel and forefoot. There was an interac-
tions, TS drop was 7.6 mm greater than MS. The overall tion effect, which indicated that EA decreased at a
change in drop throughout testing (not Pre-MA vs Post- faster rate for TS and forefoot. Hysteresis curves used
MA) was equivalent for both shoe conditions (0.5 mm). to calculate EA were visibly different among conditions
As a function of aging distance, ND and EA (Figure 6).
decreased (Figure 5). Comparing across shoe-thickness As a function of aging distance, %EA decreased by
conditions, TS ND was 2.4 mm greater than MS and 44% (Figure 5). Comparing across shoe-thickness con-
heel was 1.6 mm greater than forefoot. There was an ditions, %EA was 2.5% greater for TS (than MS) and
interaction effect, which indicated that ND decreased 13% greater for heel (than forefoot). There was an
at a faster rate for MS and forefoot. EA also decreased interaction effect which indicated that %EA decreased
354 mJ as a function of aging distance (Figure 5). at a faster rate for MS and forefoot.
6 Proc IMechE Part P: J Sports Engineering and Technology 00(0)

Table 3. Group mean (SD) of average behavior throughout aging and changes from 0 to 42 km (D) of dependent (i.e. output)
variables across shoe and shoe region variables derived from MA output.

MA output variables MS forefoot MS heel TS forefoot TS heel

Mean aging values


Thickness [mm] 11.1 (0.7) 16.1 (0.8) 15.7 (0.7) 28.3 (1.1)
Drop [mm]a 5.0 (0.5) 12.6 (0.9)
Net displacement [mm] 2.5 (0.4) 3.2 (0.6) 4 (0.4) 6.5 (0.9)
Energy absorbed [mJ] 447 (158) 469 (163) 648 (175) 734 (221)
% Energy absorbed [%] 57 (18) 73 (12) 64 (15) 70 (13)
Change from 0 to 42 km
D Thickness 1.2 (0.3) 1.2 (0.2) 1.2 (0.4) 1.8 (0.4)
D Dropa 0.0 (0.3) 0.6 (0.6)
D ND 1.1 (0.1) 0.5 (0.1) 1.1 (0.3) 0.2 (0.1)
D EA 422 (77) 243 (90) 487 (63) 418 (135)
D %EA 53.9 (1.5) 37.6 (8.1) 47.7 (1.0) 38.9 (3.7)

MA: mechanical aging; MS: minimalist shoe; TS: traditional shoe; ND: net displacement; EA: energy absorbed.
a
Drop is the difference in thickness between MS heel–MS forefoot and TS heel–TS forefoot.

Table 4. Statistical results of MA dependent (i.e. output) variables across shoe and shoe region variables derived from MA output.

MA output variables Result DoF F p f

Shoe effects
Thickness (mm) TS . MS 1, 28 1703.44 \ 0.01 7.91
Drop (mm) TS . MS 1, 14 618.66 \ 0.01 6.74
Net displacement (mm) TS . MS 1, 28 117.45 \ 0.01 2.28
Energy absorbed (mJ) TS . MS 1, 28 26.60 \ 0.01 4.88
% Energy absorbed (%) TS . MS 1, 28 8.26 0.01 1.14
Aging effects
Thickness (mm) Decrease 5, 140 136.65 \ 0.01 2.43
Drop (mm) Decrease 5, 70 4.73 \ 0.01 1.16
Net displacement (mm) Decrease 1, 28 522.30 \ 0.01 4.43
Energy absorbed (mJ) Decrease 1, 28 520.06 \ 0.01 4.43
% Energy absorbed (%) Decrease 1, 28 2298.75 \ 0.01 9.13
Region effects
Thickness (mm) Heel . forefoot 1, 28 1871.24 \ 0.01 8.16
Net displacement (mm) Heel . forefoot 1, 28 57.77 \ 0.01 1.75
% Energy absorbed (%) Heel . forefoot 1, 28 315.23 \ 0.01 3.49
Shoe–aging interactions Greater rate decrease
Thickness (mm) TS . MS 5, 140 5.65 \ 0.01 1.10
Net displacement (mm) MS . TS 1, 28 14.54 \ 0.01 1.23
Energy absorbed (mJ) TS . MS 1, 28 12.79 \ 0.01 1.21
% Energy absorbed (%) MS . TS 1, 28 2.31 0.04 1.04
Region–aging interactions Greater rate decrease
Thickness (mm) Heel . forefoot 5, 140 4.37 0.01 1.08
Net displacement (mm) Forefoot . heel 1, 28 154.57 \ 0.01 2.56
Energy absorbed (mJ) Forefoot . heel 1, 28 13.91 \ 0.01 1.22
% Energy absorbed (%) Forefoot . heel 1, 28 54.45 \ 0.01 1.71

MA: mechanical aging; MS: minimalist shoe; TS: traditional shoe; DoF: degree of freedom; f: the standardized mean difference expressed as Cohen’s f.

Human kinematics and kinetics Comparing MS and TS shoe conditions, participants


The averages of all statistically significant dependent ran in TS with 22.5°/s greater peak knee extension velo-
variables are summarized in Table 5 and significant sta- city, 1.2 mJ/kg greater knee work, 0.04 BW lower peak
tistics are summarized in Table 6. Because participants medial GRF, and a 20.5 BW/s lower vertical loading
selected treadmill speeds of either 2.4 (n = 2) or 2.5 rate (Figure 8). As a function of aging distance, ankle
(n = 2) m/s and ran at the same speed on both visits dorsiflexion decreased 3°, knee flexion velocity
(6.8 6 0.2 min/km pace), there were no differences in decreased by 1°, and vertical loading rate increased by
mean running velocity. Sagittal plane knee and ankle 4.5 BW/s (Figure 9). There were shoe–aging interaction
rotations, velocities, and joint moments are plotted in effects for knee flexion, knee power, and knee work,
Figure 7. which indicated that TS was greater than MS before
Lippa et al. 7

Figure 4. Measured (a) thickness and (b) drop before and after each human (Pre-MA, Post-MA) and shoe (MA) testing phase.
Decreases in thickness and drop represent degradation, while increases indicate recovery. Error bars represent a 95% confidence
interval.

Figure 5. Plots of (a) net displacement, (b) energy absorbed, and (c) percent energy absorbed demonstrate that the dependent
variables decreased as a function of aging distance for each of the shoe-region conditions. Error bars represent a 95% confidence
interval.
8 Proc IMechE Part P: J Sports Engineering and Technology 00(0)

Figure 6. Hysteresis curves for (a) TS heel, (b) TS forefoot, (c) MS heel, and (d) MS forefoot. The first three cycles of aging (0 km)
and 10, 21, and 42 km are displayed. The slope of the loading portion and area of the hysteresis curves represent sample stiffness
and energy absorption, respectively. The X axis title and units for (a) and (b) are same as (c) and (d).

Table 5. Group mean (SD) of select peak kinematic and average or peak kinetic variables across shoe (TS, MS) and aging conditions
(pre-MA, post-MA) derived from human testing.

Human testing variables MSPre-MA MSPost-MA TSPre-MA TSPost-MA

Kinematic variables
Peak joint angle [°]
Knee flexion 39.6 (4) 39.7 (3.1) 41.8 (3.7) 39.8 (2.5)
Ankle dorsiflexion 19.9 (3) 16.5 (2) 20.4 (2) 17.9 (1.2)
Peak joint velocity [°/s]
Knee flexion 39.6 (4) 39.7 (3.1) 41.8 (3.7) 39.8 (2.5)
Knee extension 256 (17) 250 (19) 278 (10) 273 (22)
Kinetic variables
Peak medial GRF [BW] 0.17 (0.06) 0.17 (0.03) 0.13 (0.04) 0.13 (0.05)
Loading rate [BW/s] 53 (11) 60 (13) 35 (6) 37 (7)
Knee power [W/kg] 9.8 (3.0) 10.0 (1.4) 11.7 (2.6) 10.6 (0.9)
Knee work [mJ/kg] 7.4 (1.5) 7.6 (0.8) 9.0 (1.5) 8.4 (0.9)

MA: mechanical aging; MS: minimalist shoe; TS: traditional shoe; GRF: ground reaction force.

aging, but the shoe conditions were equivalent after Discussion


aging (Figure 10). There were no significant differences
Our goal was to determine how traditional and minim-
in the other kinematics and kinetics.
alist shoes degraded over the time-course equivalent to
Lippa et al. 9

Table 6. Statistical results of select peak kinematic and average or peak kinetic variables across shoe (TS, MS) and aging conditions
(pre-MA, post-MA) derived from human testing.

Human variables Result DoF F p f

Shoe effects
Knee extension velocity (°/s) TS . MS 1, 3 13.74 0.03 2.14
Peak medial GRF (BW) MS . TS 1, 3 12.57 0.04 2.04
Vertical loading rate (BW/s) MS . TS 1, 3 17.09 0.03 2.39
Knee work (mJ/kg) TS . MS 1, 3 15.84 0.03 2.30
Aging effects
Peak ankle dorsiflexion Decreased 1, 3 34.45 0.01 3.39
Knee flexion velocity (°/s) Decreased 1, 3 43.26 0.01 3.79
Vertical loading rate (BW/s) Increased 1, 3 27.88 0.01 3.05
Shoe–aging interactions
Peak knee flexiona TS Pre . all 1, 3 15.05 0.03 2.24
Knee power (W/kg)a TS Pre . all 1, 3 11.73 0.04 1.98
Knee work (mJ/kg)a TS Pre . all 1, 3 11.26 0.04 1.94

MA: mechanical aging; MS: minimalist shoe; TS: traditional shoe; GRF: ground reaction force; DoF: degree of freedom; f: the standardized mean
difference expressed as Cohen’s f.
a
Knee interactions were due to TS Pre condition.

Figure 7. Group mean plots of (a–c) sagittal plane knee and (d–f) ankle rotations, moments, and velocities for the four shoe–aging
conditions as a function of % stance. Peak knee flexion and dorsiflexion values were statistically evaluated for each variable. The X
axis titles and units for (a)–(c) are same as (d)–(f).

a marathon and how this degradation influenced warranted because, despite the attempt to mitigate run-
human running kinematics and kinetics at the knee and ning impact using footwear,30 injury rates have
ankle. The study of human–footwear interactions was remained consistent and no epidemiological study has
10 Proc IMechE Part P: J Sports Engineering and Technology 00(0)

Figure 8. Profile plots of peak ankle and knee angles and velocities demonstrate that (a) ankle dorsiflexion decreased with aging,
(b) there was a significant shoe–aging interaction for knee flexion, (c) knee extension velocity was greater for TS than MS, and (d)
knee flexion velocity decreased with aging. Error bars represent a 95% confidence interval.

Figure 9. Profile plots of relevant kinetics variables from the force-instrumented treadmill demonstrate that (a) vertical loading
rate was greater for MS than TS and increased with aging and (b) peak medial GRF was greater for MS than TS. Error bars represent
a 95% confidence interval.

sufficiently demonstrated that shoes can reduce the rate kinetics,4–7 the MA input parameters also varied as a
of overuse running injury.3 Therefore, the first step to function of both shoe type and shoe region (Tables 1
understanding the role of shoes and injury was to and 2). Certain shoe–region combinations yielded more
understand concomitant changes in footwear material aggressive input parameters, such as (1) the forefoot
properties and athlete running biomechanics. Overall, had a higher peak force and shorter recovery time than
it was expected that degradation and biomechanics the heel, (2) the heel had greater stress and a higher
would interact in an interdependent way to produce loading rate than forefoot, and (3) there was a greater
consistent performance and running differences number of cycles and higher loading rate in the MS
between traditional and minimalist shoes. compared to the TS. Heel and forefoot appeared to
have ‘‘degradation tradeoffs’’ which prevented any pre-
diction of which region would deteriorate more rapidly.
MA input parameters Conversely, MS was expected to deteriorate more rap-
The MA input parameters were not estimates or repre- idly than TS because all MS input parameters were
sentative impulses,15–22 but were actual unique impulses more aggressive (Tables 1 and 2). While the partici-
measured for each participant during the first running pants were expected to run differently in each shoe
session (Figure 3). Because each of the unaged shoe type, this study improved understanding of the conse-
conditions demonstrated different kinematics and quences of footwear selection: differences in shoe type
Lippa et al. 11

Figure 10. Profile plots of (a–c) stiffness, power, and work for knee flexion and (d–f) ankle dorsiflexion in the landing phase of
stance. There were shoe–aging interactions for (b) knee power and (c) knee work. There were no differences in knee stiffness or
ankle variables. Error bars represent a 95% confidence interval.

affected participant kinematics, which in turn resulted Footwear degradation


in unique footwear degradation patterns. Footwear degradation was exhibited by decreases in
Similarly, input parameters were expected to affect
thickness, ND, EA, and %EA with aging distance
the level of material degradation because degradation (Figures 4 and 5) and indicated that the cell walls
by cyclic fatigue has been shown to be quite sensitive to underwent bending, breaking, and permanent plastic
selected input parameters.22,42 Degradation has been deformation,1,44 characteristic of cell flattening, foam
shown to proceed more rapidly due to several factors densification, and reduced performance.18,21,29 Material
related to running kinematics and kinetics, including stiffness also increased with aging, in agreement with
(1) decreasing the recovery time between impacts42 or previous literature.25,45–47 The loading region of the
(2) increasing the number of cycles, peak force, engi- hysteresis curves did not show a yield point (Figure 6),
neering stress, or loading rate.42,43 However, under- which supported previous reports that the microscopic
standing the origin and degree of footwear degradation cell face buckling characteristic of polymer foams1,48–50
was improved by also quantifying the subject-driven
was masked by rubber outsole deformation.51
input parameters (i.e. specific variables directly related Certain differences in compressive behavior for each
to kinematics and kinetics), which helped explain why shoe-region combination were attributed to initial mid-
degradation generally occurred more rapidly for the sole foam thickness.14 For example, TS and heel under-
MS shoe and forefoot region, compared to TS and heel, went more displacement and absorbed more energy
respectively.
12 Proc IMechE Part P: J Sports Engineering and Technology 00(0)

than MS and forefoot on average because the thicker The current results suggested that participants did
foams were more compliant (Figure 6), which resulted not substantially adapt their running patterns to a
in a less aggressive loading rate, greater deformation, more distal footstrike in MS. A substantial change in
and larger hysteresis area (Figures 3, 5 and 6). footstrike pattern between the TS and MS was not
Conversely, other degradation patterns were largely expected because all participants demonstrated a heel-
influenced by participants’ kinematics and kinetics and strike pattern during the pre-test running trial and were
were counter to what one would expect based on thick- not instructed to change their natural, habitual foot-
ness alone. From a materials perspective, thicker foams strike during testing. Inspection of the vertical GRF
were expected14 to undergo more rapid deterioration profiles showed that all running tests in both shoes
due to (1) greater compliance, (2) a greater viscous demonstrated both a passive impact peak and an active
response,1,21,52 (3) greater strain recovery rates needed propulsion peak, indicative of a heelstrike pattern.
to return to 0 mm displacement, and (4) lower foam Average vertical loading rates were, however, greater in
relative density.48,53,54 However, there were several the MS compared to the TS.6 The thinner MS heel
instances where thinner foams degraded faster than foam was stiffer (Figure 6) and provided less compres-
thicker foams (Table 4), which was due to the unique sion for deceleration during heelstrike. Previous studies
running biomechanics in each shoe condition. also found that participants wearing MS maintained a
Participants ran with higher peak forces, less recovery heelstrike and exhibited only slight (if any) kinematics
time on the forefoot region, a greater stride frequency changes such that impulse curves were bimodal and
(i.e. number of cycles), and a higher loading rate in MS loading rates and peak forces increased.2,5,7,8 While
shoes. These conditions resulted in the forefoot and these differences in TS and MS were expected based on
MS degrading more rapidly than the heel and TS, previous literature on minimalist footwear,2,5,7,8 the
respectively. Results confirmed that the difference in effects of footwear degradation warrant further
TS and MS degradation rates were due to the input discussion.
parameters (based on human kinematics and kinetics) Knee flexion velocity decreased with aging and sev-
rather than thickness or inherent material properties. eral aging–shoe interactions were observed: (1) peak
Although ND, EA, and %EA decreased more rap- knee flexion, (2) knee power, and (3) total work at the
idly for thinner-foamed shoes/regions, thickness knee in the absorption phase of stance. The results are
decreased more for thicker heel than forefoot, which supported by previous literature that reported cush-
resulted in an overall decrease in drop.14 A change in ioned shoes to elicit higher peak knee flexion compared
drop due to MA was only evident for TS and was small to non-cushioned shoes.11,57 Furthermore, the current
(D = 0.6 mm). A similar comparison of thickness results demonstrated that reducing cushioning via MA
immediately before and after human participants ran in reduced peak knee flexion velocity, power, and work.
the shoes suggested that drop did not significantly Because participants maintained a heelstrike gait in all
change. conditions, we posit that the graded heel of TS-Pre
In summary, some results were expected due to caused participants to run with greater peak knee flex-
material differences, while other findings deviated from ion and ankle dorsiflexion than when running in an
previous reports due to the incorporation of real human MS. After aging, the overall midsole thickness, energy
data rather than uniform prescribed input parameters. absorption, and drop of the shoe decreased. As a result,
As expected, thicker foams were more compliant and TS-Post mid-stance knee flexion and dorsiflexion
underwent more displacement, such that more energy decreased, which consequently affected knee power and
was absorbed and in certain ways more degradation total work.
occurred. However, thinner foams degraded more rap- Interpreting the findings presented herein, one can
idly not because of material differences, but because calculate several interesting relationships for the habi-
they were subjected to more aggressive input para- tually shod heelstrike runners as a function of footwear
meters due to how participants ran in the shoes. type. For the MS, a 243 mJ (38%) decrease in energy in
the shoe’s heel and 422 mJ (54%) in the forefoot led to
(1) 3.4° decrease in peak ankle dorsiflexion and (2)
Human kinematics and kinetics 7.0 BW/s increased loading rate. For the TS, a 418 mJ
Kinematics and kinetics were affected by wearing tradi- (39%) decrease in energy in the shoe’s heel and 487 mJ
tional versus minimalist shoes, which is consistent with (48%) in the forefoot led to (1) 2.5° decrease in peak
the literature.4–6, 9–13 Although a heelstrike gait was ankle dorsiflexion, (2) 2°/s decrease in peak knee flex-
observed for both MS and TS conditions, knee exten- ion velocity, and (3) 2.0 BW/s increased loading rate.
sion velocity was significantly higher for TS and non-
significant trends were observed (i.e. lower stride
frequency, higher dorsiflexion) which suggested that Limitations
the participants potentially ran with longer strides in A limitation of this study was that the outsole geometry
the TS.12,13 Vertical loading rate and peak medial GRF and shoe design were different between shoe condi-
were also higher when participants wore MS. Both fac- tions. Apart from customizing footwear samples, which
tors have been suggested to increase injury risk.55,56 would defeat the applicability of this work to
Lippa et al. 13

commercial running shoes and injury, a thorough References


review of available options revealed these two condi- 1. Mills N. Polymer foams handbook. London: Butterworth-
tions to be the best approximation of similarity. Heinemann, 2007.
A second limitation was that the actual shoe thick- 2. Heidenfelder J, Sterzing T and Milani TL. Systematically
ness could not be measured during human testing and modified crash-pad reduces impact shock in running
needed to be estimated. Because kinematic and kinetic shoes. Foot Sci 2010; 2: 85–91.
data were captured in the last 30 s of the subject run- 3. Richards CE, Magin PJ and Callister R. Is your prescrip-
ning trials (last 30% of total running time), the best tion of distance running shoes evidence-based? Br J
thickness estimates were after baseline and follow-up Sports Med 2009; 43: 159–162.
4. Moore IS, Pitt W, Nunns M, et al. Effects of a seven-
running test because thickness was measured immedi-
week minimalist footwear transition programme on foot-
ately after the participants removed their shoes. strike modality, pressure variables and loading rates.
However, thickness and drop were likely dynamic Foot Sci 2014; 7: 17–29.
properties during each running session. 5. Hamill J, Russell EM, Gruber AH, et al. Impact charac-
A third limitation was that only knee and ankle teristics in shod and barefoot running. Foot Sci 2011; 3:
sagittal plane kinematics and kinetics were reported. 33–40.
Therefore, the intuitive assumption that human- 6. Lieberman DE, Venkadesan M, Werbel WA, et al. Foot
generated energy absorption would increase in response strike patterns and collision forces in habitually barefoot
to shoe wear was an oversimplification. The mechanical versus shod runners. Nature 2010; 463: 531–535.
work of different body segments, movements in frontal 7. Horvais N and Samozino P. Effect of midsole geometry
and transverse planes, and metabolic cost must be con- on foot-strike pattern and running kinematics. Foot Sci
2013; 5: 81–89.
sidered in future approaches.
8. Deneweth JM, McGinnis R, Zernicke R, et al. Individ-
ual-specific determinants of successful adaptation to min-
imal and maximal running shoes. Foot Sci 2015; 7: S97–
Conclusion S99.
In summary, aging the footwear decreased the midsole 9. Moore IS, Jones A and Dixon S. The pursuit of
improved running performance: can changes in cushion-
thickness, increased the material stiffness, and reduced
ing and somatosensory feedback influence running econ-
the energy absorption of the shoe. Thinner foams, such omy and injury risk? Foot Sci 2014; 6: 1–11.
as that in the forefoot and in the MS, degraded more 10. Divert C, Mornieux G, Baur H, et al. Mechanical com-
rapidly than thicker foams. These changes in footwear parison of barefoot and shod running. Int J Sports Med
properties were associated with a decrease in ankle dor- 2005; 26: 593–598.
siflexion during running, which may have been a 11. Bonacci J, Vicenzino B, Spratford W, et al. Take your
response to the reduced attenuation of impact forces. shoes off to reduce patellofemoral joint stress during run-
Despite the changes in kinematics, greater vertical load- ning. Br J Sports Med 2014; 48: 425–428.
ing rates were observed when running in the aged 12. Squadrone R and Gallozzi C. Biomechanical and physio-
shoes. logical comparison of barefoot and two shod conditions
in experienced barefoot runners. J Sports Med Phys Fit
2009; 49: 6–13.
Declaration of conflicting interests 13. Franz JR, Wierzbinski CM and Kram R. Metabolic cost
The author(s) declared no potential conflicts of interest of running barefoot versus shod: is lighter better. Med Sci
Sports Exerc 2012; 44: 1519–1525.
with respect to the research, authorship, and/or publi-
14. Lippa NM, Collins PK, Bonacci J, et al. Mechanical age-
cation of this article.
ing performance of minimalist and traditional footwear
foams. Foot Sci 2017; 9: 9–20.
Funding 15. Cook SD, Kester MA and Brunet ME. Shock absorption
characteristics of running shoes. Am J Sports Med 1985;
The author(s) disclosed receipt of the following finan-
13: 248–253.
cial support for the research, authorship, and/or publi- 16. Brückner K, Odenwald S, Schwanitz S, et al. Polyur-
cation of this article: Doctoral student support was ethane-foam midsoles in running shoes—impact energy
provided by the Australian Government via the and damping. Procedia Eng 2010; 2: 2789–2793.
Endeavour Fellowship [grant number 4623-2015]. The 17. Schwanitz S, Moser S and Odenwald S. Comparison of
Applied Impact Biomechanics Lab in the School of test methods to quantify shock attenuating properties of
Kinesiology and Nutrition is supported by the athletic footwear. Procedia Eng 2010; 2: 2805–2810.
Department of Defense and U.S. Army Award 18. Schwanitz S and Odenwald S. Long-term cushioning
#NSRDEC-W911QY-15-C-0038. properties of running shoes. In: Brisson P (ed.) The
engineering of sport 7. New York: Springer, 2008,
pp.95–100.
ORCID iD 19. Wang L, Xian Li J, Hong Y, et al. Changes in heel cush-
Trenton E Gould https://orcid.org/0000-0002-5905- ioning characteristics of running shoes with running mile-
2206 age. Foot Sci 2010; 2: 141–147.
14 Proc IMechE Part P: J Sports Engineering and Technology 00(0)

20. Dib M, Smith J, Bernhardt K, et al. Effect of environ- 39. van denBogert AJ, Smith GD and Nigg BM. In vivo
mental temperature on shock absorption properties of determination of the anatomical axes of the ankle joint
running shoes. Clin J Sport Med 2005; 15: 172–176. complex: an optimization approach. J Biomech 1994; 27:
21. Sun P-C, Wei H-W, Chen C-H, et al. Effects of varying 1477–1488.
material properties on the load deformation characteris- 40. Moisio KC, Sumner DR, Shott S, et al. Normalization of
tics of heel cushions. Med Eng Phys 2008; 30: 687–692. joint moments during gait: a comparison of two tech-
22. Lippa N, Hall E, Piland S, et al. Mechanical ageing pro- niques. J Biomech 2003; 36: 599–603.
tocol selection affects macroscopic performance and 41. Masson M and Loftus G. Using confidence intervals for
molecular level properties of Ethylene Vinyl Acetate graphically based interpretation. Can J Exp Psychol 2003;
(EVA) running shoe midsole foam. Procedia Eng 2014; 57: 203–220.
72: 285–291. 42. Lippa NM, Krzeminski DE, Piland SG, et al. Biofidelic
23. Taunton JE, Ryan MB, Clement DB, et al. A prospective mechanical ageing of ethylene vinyl acetate running foot-
study of running injuries: the Vancouver sun run ‘‘in wear midsole foam. Proc IMechE, Part P: J Sports Engi-
training’’ clinics. Br J Sports Med 2003; 37: 239–244. neering and Technology 2016; 231: 287–297.
24. Kong PW, Candelaria NG and Smith DR. Running in 43. Ouellet S, Cronin D and Worswick M. Compressive
new and worn shoes: a comparison of three types of cush- response of polymeric foams under quasi-static, medium
ioning footwear. Br J Sports Med 2009; 43: 745–749. and high strain rate conditions. Polym Test 2006; 25:
25. O’Leary K, Vorpahl KA and Heiderscheit B. Effect of 731–743.
cushioned insoles on impact forces during running. J Am 44. Mills NJ and Rodriguez-Perez MA. Modelling the gas-
Podiatr Med Assoc 2008; 98: 36–41. loss creep mechanism in EVA foam from running shoes.
26. Hreljac A. Impact and overuse injuries in runners. Med Cell Polym 2001; 20: 79–100.
Sci Sports Exerc 2004; 36: 845–849. 45. Mills C, Yeadon MR and Pain MT. Modifying landing
27. Hreljac A, Marshall RN and Hume PA. Evaluation of mat material properties may decrease peak contact forces
lower extremity overuse injury potential in runners. Med but increase forefoot forces in gymnastics landings.
Sci Sports Exerc 2000; 32: 1635–1641. Sports Biomech 2010; 9: 153–164.
28. Chambon N, Sevrez V, Ly QH, et al. Aging of running 46. Even-Tzur N, Weisz E, Hirsch-Falk Y, et al. Role of
shoes and its effect on mechanical and biomechanical EVA viscoelastic properties in the protective performance
variables: implications for runners. J Sports Sci 2014; 32: of a sport shoe: computational studies. Biomed Mater
1013–1022. Eng 2006; 16: 289–299.
29. Verdejo R and Mills NJ. Simulating the effects of long 47. Benanti M, Andena L, Briatico-Vangosa F, et al. Viscoe-
distance running on shoe midsole foam. Polym Test 2004; lastic behavior of athletics track surfaces in relation to
23: 567–574. their force reduction. Polym Test 2013; 32: 52–59.
30. Shorten M. Running shoe design: protection and perfor- 48. Gibson LJ and Ashby MF. Cellular solids. 2nd ed. Cam-
mance. In: Pedoe DT (ed.) Marathon medicine. London: bridge: Cambridge University Press, 1997.
The Royal Society of Medicine Press, 2000, pp.159–169. 49. Ashby MF. The mechanical properties of cellular solids.
31. Taunton JE, Ryan MB, Clement DB, et al. A retrospec- Metall Trans A 1983; 14A: 1755–1769.
tive case-control analysis of 2002 running injuries. Br J 50. Mills NJ and Zhu HX. The high strain compression of
Sports Med 2002; 36: 95–101. closed-cell polymer foams. J Mech Phys Solids 1999; 47:
32. Gardner LI, Dziados JE, Jones BH, et al. Prevention of 669–695.
lower extremity stress fractures: a controlled trial of a 51. Silva RM, Rodrigues JL, Pinto VV, et al. Evaluation of
shock absorbent insole. Am J Public Health 1988; 78: shock absorption properties of rubber materials regard-
1563–1567. ing footwear applications. Polym Test 2009; 28: 642–647.
33. McLean SG, Su A and van den Bogert AJ. Development 52. Joubert C, Michel A, Choplin L, et al. Influence of the
and validation of a 3-D model to predict knee joint load- crosslink network structure on stress-relaxation behavior:
ing during dynamic movement. J Biomech Eng 2003; 125:
viscoelastic modeling of the compression set experiment.
864–874.
J Polym Sci, Part B: Polym Phys 2003; 41: 1779–1790.
34. Divert C, Baur H, Mornieux G, et al. Stiffness adapta-
53. Han CD and Yoo HJ. Studies on structural foam pro-
tions in shod running. J Appl Biomech 2005; 21: 311–321.
cessing. IV. Bubble growth during mold filling. Polym
35. Leardini A, Chiari L, Croce UD, et al. Human movement
Eng Sci 1981; 21: 518–533.
analysis using stereophotogrammetry: part 3. Soft tissue
54. Liao R, Yu W and Zhou C. Rheological control in foam-
artifact assessment and compensation. Gait Posture 2005;
ing polymeric materials: II. Semi-crystalline polymers.
21: 212–225.
Polymer 2010; 51: 6334–6345.
36. McLean SG, Borotikar B and Lucey SM. Lower limb
55. Cheung RT, Wong MY and Ng GY. Effects of motion
muscle pre-motor time measures during a choice reaction
control footwear on running: a systematic review. J
task associate with knee abduction loads during dynamic
Sports Sci 2011; 29: 1311–1319.
single leg landings. Clin Biomech 2010; 25: 563–569.
56. Daoud AI, Geissler GJ, Wang F, et al. Foot strike and
37. McLean SG and Samorezov JE. Fatigue-induced ACL
injury rates in endurance runners: a retrospective study.
injury risk stems from a degradation in central control.
Med Sci Sports Exerc 2012; 44: 1325–1334.
Med Sci Sports Exerc 2009; 41: 1661–1672.
57. De Wit B, De Clercq D and Aerts P. Biomechanical anal-
38. Bell AL, Pedersen DR and Brand RA. A comparison of
ysis of the stance phase during barefoot and shod run-
the accuracy of several hip center location prediction
ning. J Biomech 2000; 33: 269–278.
methods. J Biomech 1990; 23: 617–621.

You might also like