throbber
ARTICLE IN PRESS
`
`Journal of Biomechanics 37 (2004) 1379–1386
`
`Heel–shoe interactions and the durability of EVA foam
`running-shoe midsoles
`
`R. Verdejo*, N.J. Mills
`
`Metallurgy and Materials, University of Birmingham, Birmingham B15 2TT, UK
`
`Accepted 16 December 2003
`
`Abstract
`
`A finite element analysis (FEA) was made of the stress distribution in the heelpad and a running shoe midsole, using heelpad
`properties deduced from published force-deflection data, and measured foam properties. The heelpad has a lower initial shear
`modulus than the foam (100 vs. 1050 kPa), but a higher bulk modulus. The heelpad is more non-linear, with a higher Ogden strain
`energy function exponent than the foam (30 vs. 4). Measurements of plantar pressure distribution in running shoes confirmed the
`FEA. The peak plantar pressure increased on average by 100% after 500 km run. Scanning electron microscopy shows that
`structural damage (wrinkling of faces and some holes) occurred in the foam after 750 km run. Fatigue of the foam reduces heelstrike
`cushioning, and is a possible cause of running injuries.
`r 2004 Elsevier Ltd. All rights reserved.
`
`Keywords: FEA; Foam; Heelpad; Running; Plantar pressure distribution
`
`1. Introduction
`
`Running involves a series of heel-strikes on the
`ground. The midsole foams of running shoes, by
`absorbing energy, limit the peak impact force in the
`heel-strike. Shorten (2000) showed that soft cushioning
`systems increased the duration of footstrike impacts and
`spread the load across a larger area of the plantar surface.
`Plantar surface pressure distributions were measured for
`running in various shoe types by Chen et al. (1994),
`Henning and Milani (1995, 2000), and Shiang (1997).
`However, there is no published information on how the
`pressure distribution changes with shoe use.
`The impact response of a heel on a flat rigid surface
`was shown to be non-linear, with energy absorption on
`unloading, by Cavanagh et al. (1984). In these experi-
`ments the heelpad deformation is overestimated, since
`the lower leg and knee deform to some extent. Aerts
`et al. (1995, 1996) measured the impact response of the
`isolated lower half of the foot, giving a better indication
`of the heelpad response. However, no one has simulta-
`
`*Corresponding author. Tel.: +44-121-414-5180; fax: +44-121-414-
`5232.
`E-mail addresses: raquel@verdejo.net (R. Verdejo),
`n.j.mills@bham.ac.uk (N.J. Mills).
`
`0021-9290/$ - see front matter r 2004 Elsevier Ltd. All rights reserved.
`doi:10.1016/j.jbiomech.2003.12.022
`
`neously measured the heelpad response and the pressure
`distribution at the foot/floor interface. Miller-Young
`et al. (2002) characterised the fat pad, taken from
`cadavers of the elderly, in compression, to generate data
`for finite element analysis (FEA). However the magni-
`tudes (0.01–0.1 Pa) of
`the moduli reported, appear
`unrealistically small. Aerts and De Clercq (1993)
`performed pendulum impact tests on shod heels, and
`showed that the heelpad compression was smaller with a
`harder shoe midsole; they reasoned that the heelpad
`response was rate dependent, and that the shoe heel
`counter constrained the heelpad. Gefen et al. (2001)
`measured the heelpad thickness of two 30-year old
`subjects as 11 and 13 mm. They found that the non-
`linearly elastic heelpad had an initial compressive
`modulus of 105711 kPa.
`The only detailed FEA of a foot–shoe interaction
`(Lemmon et al., 1996) was a two-dimensional analysis of
`the forefoot region for walking. Shiang (1997) per-
`formed FEA of polyurethane foam midsoles in the heel
`region, but gave no material parameters. Rather than
`model the heelpad, he used in-shoe pressure data as
`vertical loads for the upper surface of the midsole; this
`ignores shear stresses at the interface. He predicted a
`greater (mean) vertical strain in the centre of the heel
`contact area than elsewhere in the foam.
`
`MacNeil Exhibit 2179
`Yita v. MacNeil IP, IPR2020-01139, Page 1
`
`

`

`1380
`
`R. Verdejo, N.J. Mills / Journal of Biomechanics 37 (2004) 1379–1386
`
`ARTICLE IN PRESS
`
`Most running shoe midsoles are made from the
`foamed copolymer of
`ethylene and vinyl acetate
`(EVA), of density in the range 150–250 kg/m3. Foot-
`strikes, repeated at approximately 1.5 Hz, may cause
`fatigue damage to the foam, hence may lead to the foam
`bottoming out, causing injuries. Prior research has used
`laboratory fatigue tests, which only provides indirect
`evidence of performance deterioration in running.
`Misevich and Cavenagh (1984) used repeated, rapid,
`uniaxial compression tests on EVA foam, to show that
`the midsole force-deflection response changed with cycle
`number. No details of the EVA foam densities were
`given. Cook et al. (1985) used a prosthetic foot, tilted
`back by 15, to load the heel of the shoe from 0 to 1.5 kN
`at 2.5 Hz. After the equivalent of 500 miles running, the
`shoes had 55710% of their initial energy absorption.
`Bartlett (1999) discussed the cell geometry in sectioned
`EVA midsoles, claiming that cells next to the outsole
`became flattened after 3200 km of running. Mills and
`Rodriguez-Perez (2001) studied diffusion in EVA foam
`under creep loading, concluding that the air content of
`the foam cells decreased, reducing the cushioning.
`One objective was to study the mechanical interaction
`of the heelpad with running shoe midsoles, and to
`estimate the magnitude of internal heelpad stresses.
`Another was to clarify the mechanism of shoe midsole
`degradation, and to investigate the resulting changes in
`the peak plantar pressures.
`The approach taken was to use FEA of the heelpad
`and shoe to predict the pressure distribution at the heel/
`shoe interface,
`then use the pressure distribution
`experiments on runners to validate the analysis, and to
`follow changes in shoe cushioning.
`
`2. Methods
`
`2.1. Plantar pressure distribution
`
`Three healthy male long distance runners (Table 1) ran
`at 2.61 m/s for 10 min on a Quinton Instrument Co. 640
`treadmill; this short experiment avoided fatigue, hence
`possible changes in foot loading (Edington et al., 1990).
`The treadmill provides a standard running speed and
`running surface. They were all rearfoot strikers, did not
`use orthotics, and reported no lower extremity injury for
`the past year. The University ethical committee ap-
`proved the study and informed consent was obtained
`from the runners. They wore Reebok Aztrek DMX
`shoes that were new at the start of the experiment.
`Their plantar pressure distribution was recorded using
`the Tekscan ‘F-Scan’ system—a flexible, 0.18 mm thick,
`plastic sole-shape having 960 pressure sensors with
`spatial resolution of 5 mm. The resistance of pressure-
`sensitive ink, contained between two polymer-film
`substrates, decreases as the pressure, applied normal to
`
`Table 1
`Runner characteristics
`
`Run distance
`(km)
`
`Age
`
`Weight (kg)
`
`Shoe size
`(UK)
`
`Runner 1
`Runner 2
`Runner 3
`
`725
`700
`500
`
`34
`37
`49
`
`91.472.0
`82.171.4
`62.972.3
`
`11
`9
`8
`
`Fig. 1. Calibration of five new F-scan sensors, ‘non-averaged’.
`
`the substrate, increases. Ahroni et al. (1998) and Mueller
`and Strube (1996) reviewed studies on F-scan sensors;
`some researchers reported good reliability and reprodu-
`cibility, while others reported a decrease of sensor
`output with time at fixed pressure. Woodburn and
`Helliwell (1996) concluded that they were not suitable
`for accurate, repeatable measurements. However, peak
`pressure measurements in this paper were similar to
`those reported previously, allowing for differences in
`running speeds (Gross and Bunch, 1989).
`To check the sensor linearity, an area of 0.0079 m2 in
`the forefoot region was sandwiched between two 6 mm
`layers of soft, closed-cell ‘Airex 5230’ PVC foam, then
`loaded in uniaxial compression between metal plates
`using an Instron machine. Constant pressures of 31, 36,
`39, 155, 225, 329, 398 kPa were applied for 20 s, while
`data was recorded for 2 s. The loading cycle was
`repeated five times, on five new and three used sensors.
`Values were taken from 15 different frames of each 300
`frames ‘movie’. The manufacturer suggested using the
`‘weighted averaging’
`function to reduce cell-to-cell
`variation. There
`is a linear
`relationship between
`pressures measured by the F-scan sensor and those
`applied by the Instron (Fig. 1). Table 2 gives the values
`of the intercept a; the slope b and the correlation
`coefficient R2: Calibrations using ‘weighted averaging’
`and ‘non-weighted averaging’ are not
`significantly
`different. However when sensors were recalibrated, at
`
`MacNeil Exhibit 2179
`Yita v. MacNeil IP, IPR2020-01139, Page 2
`
`

`

`ARTICLE IN PRESS
`
`R. Verdejo, N.J. Mills / Journal of Biomechanics 37 (2004) 1379–1386
`
`1381
`
`tion was simplified as a hemisphere of radius 15 mm,
`attached to the end of a 20 mm long vertical cylinder of
`radius 15 mm (Fig. 8a). The heelpad outer geometry was
`taken to be a vertical cylinder of radius 30 mm; the lower
`surface was spherical with a radius of curvature of
`40 mm, typical of a foot; a smooth blend was made
`between this surface and the vertical cylindrical surface.
`The minimum heelpad thickness was taken as 12 mm,
`the mean of the values given by Gefen et al. (2001). The
`lower ends of the tibia and fibula are approximated as
`an annular projection on the upper part of
`the
`calcaneus. When a EVA midsole foam was present, it
`was taken as vertical cylinder of radius 35 mm and
`height 22 mm, with initially flat upper and lower faces.
`Its lower surface (or that of the heelpad when no foam
`was present) rested on a flat rigid support table.
`The support
`table was fixed, while the upper
`calcaneus boundary was
`ramped down by 20 mm
`(12 mm when no foam was present). The heelpad is
`allowed to expand freely at the sides; it is assumed that a
`shoe heel counter and upper would have no confining
`effect. The heel pad is assumed to be bonded to the
`calcaneus surface, which allows load transfer by shear
`stresses at
`the interface. The coefficient of
`friction
`between the heelpad and the foam or support table
`was taken as 1.0. Meshing was chosen to maximise the
`computation stability.
`
`3.2. Material models
`
`Table 2
`Tekscan sensor calibration
`
`Sensor
`age
`
`New
`New
`Used
`
`Data treatment
`
`Intercept
`a (kPa)
`
`Slope b
`
`Correlation,
`r2
`
`Non-average
`Average
`Non-average
`
`1.876
`1.894
`2.591
`
`1.023
`1.028
`1.136
`
`0.998
`0.998
`0.996
`
`Average values for five new and three used sensors.
`
`the end of their use, the slope was 11% larger than that
`for new sensors.
`The subjects were asked to run on hard surfaces (track
`or roads), and to keep a running distance diary. Every
`15 days the plantar pressure distribution was measured,
`after a standardised 24 h recovery time since the last run.
`The insole was trimmed to fit the subjects’ right shoe.
`The sensors were calibrated every session by the known
`weight of the test subject, standing on one foot. Data
`were recorded, at the beginning (once they had acquired
`their gait) middle and end of the 10 min run, at 150 Hz
`for 4 s, which gave an average of 5 strikes per movie.
`The peak pressures from each frame were corrected to
`an ‘Instron’ pressure value using the new ‘non-averaged’
`calibration of Fig. 1.
`
`2.2. Shoe midsole foam characterisation
`
`The midsole of Reebok Aztrek DMX shoes was
`20 mm thick and contained two foams: a section of area
`20 mm by 40 mm on the lateral side of the heel was
`coloured grey, and the rest white. The densities were
`measured using a hydrostatic balance. Differential
`scanning calorimetry (DSC) was carried out using a
`Mettler DSC 30. The melting point and crystallinity
`were taken from the second heating run; this eliminates
`the effect of thermal ageing. The melting point was
`taken as the peak of the heat flow vs. temperature curve;
`the degree of crystallinity was calculated by dividing the
`melting peak area by the 286.8 J/g enthalpy of fusion for
`polyethylene crystals (Brandrup and Immergut, 1975).
`Cross-sections of the midsoles, from new and used
`trainers, were fractured after
`immersion in liquid
`nitrogen, then vacuum coated with gold. They were
`examined using a JEOL JSM 5410 scanning electron
`microscopy (SEM).
`
`3. FEA
`
`3.1. Geometry and boundary conditions
`
`ABAQUS Standard FEA version 6.3 (HKS) was used
`with the large deformation option. The calcaneus
`geometry was simplified to have a vertical axis of
`rotational symmetry. The geometry of its lower projec-
`
`Ideally, energy losses in the heelpad and shoe should
`be considered. Linear viscoelasticity can be incorporated
`in Explicit FEA, but the heelpad and the foam are non-
`linear viscoelastic materials. Initially a hyperelastic
`model was used, allowing the use of the more-stable
`Implicit FEA. The hyperfoam model in ABAQUS uses
`the Ogden strain energy function:
`2 þ laiðlai1 þ lai 3 3Þ þ 1
`bi
`
`
`
`
`
`
`
`Jaibi 1
`
`;
`
`ð1Þ
`
`2mi
`a2
`i¼1
`i
`where lI are the principal extension ratios, J ¼ l1l2l3 is
`a measure of the relative volume, the mi are shear
`moduli, N is an integer, and ai and bi non-integral
`exponents. The latter are related to Poisson’s ratio ni by
`ni
`1 2ni
`X
`The initial shear modulus is given by
`
`
`
`
`
`
`
`
`
`XN
`
`U ¼
`
`bi ¼
`
`:
`
`ð2Þ
`
`m ¼
`
`mi:
`
`i¼1;N
`
`ð3Þ
`
`One of the hyperelastic models in ABAQUS is the
`Ogden strain energy function:
`ðlai1 þ lai2 þ lai3 3Þ þ 1
`Di
`
`
`
`
`
`
`
`
`

`
`J 1
`
`Þ2i
`
`;
`
`ð4Þ
`
`
`
`2mi
`a2
`i
`
`XN
`
`i¼1
`
`U ¼
`
`MacNeil Exhibit 2179
`Yita v. MacNeil IP, IPR2020-01139, Page 3
`
`

`

`1382
`
`R. Verdejo, N.J. Mills / Journal of Biomechanics 37 (2004) 1379–1386
`
`ARTICLE IN PRESS
`
`-20
`
`-10
`tensile strain %
`
`0
`
`10
`
`Fig. 3. Runner 3 pressure distribution after various distances run.
`
`400
`
`200
`
`0
`
`-200
`
`tensile stresskPa
`
`-400
`
`-30
`
`Fig. 2. Impact compression, and tensile stress–strain response for
`loading and unloading EVA foam of density 170 kg m3, compared
`with hyperelastic model prediction (dashed curve).
`
`:
`
`ð5Þ
`
`where the bulk modulus K is given by
`K ¼ 1
`Di
`Lemmon et al. (1996) used a N ¼ 3 hyperfoam model
`to fit compressive data for an EVA foam (‘Cloud’ from
`Soletech); the initial shear modulus was 479 kPa, while
`their parameters predict that the Young’s modulus in
`tension is much higher than that
`in compression.
`Soletech says the Cloud’ density is between 150 and
`200 kg m3 (personal communication).
`The heelpad was simulated using the Ogden hyper-
`elastic material, with initial compressive modulus close
`to that found by Gefen et al. (2001), and bulk modulus
`that of water (2 GPa). The non-linearity parameters
`were found by matching the force-deflection response of
`the isolated heel of a 24-year old male (‘Foot II’ of Aerts
`et al. (1996)). The response of the EVA foam from the
`Reebok shoes was measured in both uniaxial compres-
`sion and in tension. Since the response alters with cycle
`number, for the first few cycles of deformation, data was
`taken after 10 cycles. Fig. 2 shows the combined
`response for tension and compression, on loading and
`unloading for the 10th cycle. The EVA foam response
`was modelled using the Ogden hyperfoam material, with
`N ¼ 2 and shear modulus m1=1000 kPa, a1 ¼ 10;
`Poisson’s ratio n1 ¼ 0; m2 ¼ 50 kPa, a2 ¼ 4; n2 ¼ 0:4:
`Fig. 2 shows that this provides a reasonable match to
`both the tensile and compressive response.
`
`4. Results
`
`4.1. Plantar pressure distribution
`
`Fig. 4. Peak pressure in runner 3’s heel region as a function of run
`distance.
`
`pressure is a maximum in the heel region. This local
`peak almost has the axial symmetry assumed in the
`FEA, although the centre of this peak is offset to the
`medial side of the foot for runner 3. The peak pressures
`are lower than some reported values (Henning and
`Milani, 1995, 2000), due to the running speed being low,
`and the surface being more compliant than a running
`track.
`The peak pressures in the heel region increased with
`running time (in the 10 min experiment), with a greater
`increase in the second and subsequent sessions, when the
`trainers had been used for several 100 km. There is an
`average 100% increase in peak pressure over the total
`run distance for all three runners (Fig. 4).
`
`4.2. Foam characterisation
`
`The Tekscan measurements of pressure distributions
`(Fig. 3) are for the time in the footstrike when the peak
`
`The density and DSC results (Table 3) are approxi-
`mately the same for both white and grey foam samples,
`suggesting that they only differ in colour. The degree of
`
`MacNeil Exhibit 2179
`Yita v. MacNeil IP, IPR2020-01139, Page 4
`
`

`

`ARTICLE IN PRESS
`
`R. Verdejo, N.J. Mills / Journal of Biomechanics 37 (2004) 1379–1386
`
`1383
`
`in the
`
`crystallinity suggests a 18% VA content
`copolymer (Dupont, 1997).
`The elastic moduli of EVA foams increases with their
`density. Larger shoe sizes may have higher foam
`densities, to give extra support to the assumed heavier
`wearers. Samples were cut from the arch region of each
`used running shoe, a region that suffers little damage.
`The size 11 shoe has increased foam density (Table 4),
`but sizes 8 and 9 the same density, allowing for density
`variation from the manufacturing process.
`
`Micrographs of the midsole foam from the new
`trainers (Fig. 5) contain flattened cells close to the lower
`(outsole) and upper (insole) surfaces; they result from
`moulding the midsole into its final shape. The cells are
`slightly elongated along the shoe length direction.
`Hence, Bartlett (1999) may have incorrectly interpreted
`flattened, near-surface cells as evidence of foam fatigue.
`All the trainers at the end of the experiment (Fig. 6)
`contain some cell faces that are wrinkled. More severe
`damage, such as cell-face fractures, is present in the
`trainers of runners 1 and 2. Fractures in cut cell faces
`may be due to sample preparation; consequently only
`fractures in complete cell faces are considered to be due
`to trainer use.
`
`The best fit to Aerts et al. (1996) data for a heel pad
`impacted by a flat rigid surface was obtained (Fig. 7)
`using shear moduli m1 ¼ m2 ¼ 50 kPa, with exponents
`a1 ¼ 30 and a2 ¼ 4: Although the material model
`cannot simulate hysteresis on unloading, it predicts the
`shape of the loading curve.
`Figs. 8b and c show the vertical compressive stress s22
`contours, respectively, for the bare heel on flat anvil and
`the heel on shoe, at total deformations of 4 and 10 mm,
`respectively, when the force is close to 0.5 kN. The total
`force on the foot will be higher, since other parts of the
`foot also transit force to the ground (Fig. 3). The main
`
`Crystallinity
`(%)
`
`4.3. FEA results
`
`Table 3
`Characterisation of EVA foams in Reebok midsole
`
`Sample Density
`(kg/m3)
`
`Melting
`point (C)
`
`White
`Grey
`
`170
`173
`
`82.2
`82.0
`
`Melting
`enthalpy
`(J/g)
`
`56.7
`53.6
`
`19.8
`18.7
`
`Table 4
`EVA foam density in used trainers
`
`Runner
`
`Runner 1
`Runner 2
`Runner 3
`Unused
`
`Shoe size
`
`Density (kg/m3)
`
`11
`9
`8
`8
`
`19073
`16875
`16474
`17073
`
`Fig. 5. SEM micrographs of EVA foam from new trainers, with the midsole thickness direction horizontal, and the shoe length direction vertical: (a)
`near the lower surface, with thick cell faces on the right (next to outsole), and (b) centre of midsole.
`
`Fig. 6. The heel region of trainers after (a) 500 km and (b) 750 km, showing some wrinkled cell faces. In (b) there are holes in some internal cell faces.
`
`MacNeil Exhibit 2179
`Yita v. MacNeil IP, IPR2020-01139, Page 5
`
`

`

`1384
`
`R. Verdejo, N.J. Mills / Journal of Biomechanics 37 (2004) 1379–1386
`
`ARTICLE IN PRESS
`
`peak force during the footstrike of runner 3 was
`typically 1.0 kN, with an initial peak of 0.8 kN. These
`compares with values of 2.3 and 1.7 kN, respectively, for
`running at 3.6 m/s on a rigid surface (Gross and Bunch,
`1989). Their peak heel pressures increased from 300 to
`420 kPa, when the running speed increased from 3.0 to
`4.5 m/s.
`the predicted
`force close to 0.5 kN,
`For a heel
`maximum compressive stress on the heelpad/foam
`interface
`is approximately 0.7 MPa, while
`in the
`bare heel test (Fig. 8b), it is 2.0 MPa. The maximum
`stress at the bone–heelpad interface is 1.0 MPa for the
`shod foot.
`In the foot/shoe simulation, for forces less than 200 N,
`the majority of the deformation occurs by the flattening
`of the lower surface of the heelpad, and the increase of
`the contact area with the foam. However, at higher
`
`forces, the deformed heelpad does not decrease much in
`thickness, while the midsole upper surface becomes
`increasingly concave. Although the heelpad has spread
`laterally, the side of the EVA foam hardly bulges. The
`maximum foam stress occurs at the centre of the contact
`area on the foam upper surface (Fig. 8c). This confirms
`the Tekscan data for the pressure map on the upper
`surface of the shoe midsole—with a 300 kPa stress area
`of diameter approximately 20 mm.
`The energy of the footstrike was calculated as the
`integral of the force vs. deflection graph, for the heel
`plus midsole simulation. Fig. 9 shows the force is a
`nearly linear function of the footstrike energy; the result
`of the force increasing nearly exponentially with the
`deflection, but not as rapidly as for the bare heel in
`Fig. 7. The maximum vertical compressive stress in the
`foam, which occurs at the centre of the heel/foam
`
`force
`no shoe
`
`stress
`
`force
`with shoe
`
`2
`
`4
`
`6
`
`8
`
`10
`
`energy J
`
`2.5
`
`2
`
`1.5
`
`1
`
`0.5
`
`peak stress(MPa) or force (kN)
`
`0
`
`0
`
`Fig. 7. Force vs. displacement, for impacts on an isolated human heel,
`redrawn from Aerts et al. (1996). FEA prediction for heel model
`(dashed curve).
`
`Fig. 9. Predicted force (and peak compressive stress on heelpad
`surface) vs. energy input for heelpad on an EVA midsole, compared
`with force vs. energy graph without a shoe.
`
`Fig. 8. (a) Undeformed mesh. Contours of vertical compressive stress (kPa) for heel on flat surface; (b) no shoe; 424 N force, 4.0 mm deflection; and
`(c) EVA foam shoe, 502 N force, 10.2 mm deflection.
`
`MacNeil Exhibit 2179
`Yita v. MacNeil IP, IPR2020-01139, Page 6
`
`

`

`ARTICLE IN PRESS
`
`R. Verdejo, N.J. Mills / Journal of Biomechanics 37 (2004) 1379–1386
`
`1385
`
`interface, is a roughly linear function of the footstrike
`energy. Hence a footstrike of a given kinetic energy will
`produce a much lower peak force for the shod foot
`rather than the bare heel.
`
`5. Discussion
`
`Both the heelpad and the running shoe EVA foam act
`as shock-absorbing non-linear spring-dashpot struc-
`tures, reducing the peak force in a heelstrike. There is
`a synergy in their responses; the foam, by indenting on
`its upper surface, increases the load spreading to the
`plantar surface, which reduces the force on the heel area.
`The indentation also probably stabilises the angular
`position of the calcaneus, affecting foot pronation. The
`properties of EVA foam can be measured more easily
`than those of the human heelpad; the modelling of their
`interaction is also a method of comparing their relative
`mechanical properties.
`FEA has successfully analysed the non-linear, large
`deformation, problem of heel/shoe interaction. The
`predicted lack of bulging of the shoe foam sides is
`consistent with experimental observations; however
`such bulging is predicted for softer, low-density EVA
`foams. The EVA foam is more compliant than the
`heelpad, since it has a much lower bulk modulus, in spite
`of its shear modulus being higher. The initial foam shear
`modulus used here is double that used by Lemmon et al.
`(1996), for an EVA foam density inside the range
`estimated for their Soletech foam. Their Ogden strain
`energy function exponent a2 ¼ 3:9 is smaller than the
`a1 ¼ 10 used here, but both values provide significant
`hardening in compression. The heelpad initial shear
`modulus used here is 100 kPa,
`the same order of
`magnitude as their value used for forefoot soft tissue,
`and equal to the value given by Gefen et al. (2001).
`Hence the data of Miller-Young et al. (2002) must be in
`error by orders of magnitude. The uncertainty in the
`moduli is probably an order of two, since the exact
`dimensions of the heel tested by Aerts et al. (1996) is
`unknown, and since viscoelasticity was ignored in the
`modelling. It appears that heelpad is highly hyperelastic,
`with a Ogden strain energy function exponent close to
`30. This is higher than the value of 17 used to model the
`thigh tissue (Setybudhy et al., 1997). The predicted
`pressure distribution at the skin/midsole interface is
`confirmed qualitatively by the F-scan data. In an ideal
`experiment the interface pressure, and the deformation
`of the heelpad and foam, would be simultaneously
`measured for a runner; however this is impossible at
`present.
`FEA predicts a significantly higher peak heelpad
`pressure in a bare heel strike, compared with a shod heel
`strike with the same force; however barefoot and shod
`runners run differently. The pressure distribution on the
`
`midsole upper surface is non-uniform, with peak values
`near the centre of the heel contact area. The impacts
`cause fatigue damage to the EVA foam. Although air
`compression provides a major shock cushioning me-
`chanism in the foam, other experiments (Verdejo and
`Mills, 2004) suggest that air loss is negligible in running.
`Instead, weakening of the EVA structure causes soft-
`ening of the foam. The wrinkling of some cell faces is
`evidence of the foam fatigue. The consequent increase in
`peak plantar pressures is a real effect, since it is an order
`of magnitude larger than the sensitivity change of
`Tekscan F-scan sensors.
`One limitation of the plantar pressure measurements
`is that they were made on a treadmill. The biomechanics
`of treadmill running differ from those of overground
`running (Nigg et al., 1995), and the peak footstrike force
`will be lower on the more-compliant treadmill surface.
`This means that the peak plantar pressures will be
`higher for running on a road than those measured for
`running on a treadmill.
`Modelling of EVA foam midsoles, with a region
`weakened by the heel-strike stress field, has not yet been
`attempted; further FEA will consider a weakened region
`in upper centre of
`the midsole. Recent
`research
`(Taunton et al., 2003) suggests that running shoe age
`contributes to running injuries. The deterioration of
`running shoe cushioning may be an important explana-
`tory factor for such an effect.
`
`Acknowledgements
`
`R. Verdejo would like to thank EPSRC for support in
`the form of a Ph.D. studentship. We would like to thank
`the runners who took part in the experiments. We would
`like to thank one referee for suggesting a better source
`for the heelpad mechanical response.
`
`References
`
`Aerts, P., Ker, R.F., et al., 1995. The mechanical properties of the
`human heel pad. Journal of Biomechanics 28, 1299–1304.
`Aerts, P., Ker, R.F., De Clercq, D., et al., 1996. The effects of isolation
`on the mechanics of the human heel pad. Journal of Anatomy 188,
`417–423.
`Aerts, P., De Clercq, D., 1993. Deformation characteristics of the heel
`region of the shod foot during a simulated heelstrike. Journal of
`Sports Science 11, 449–461.
`Ahroni, J.H., Boyco, E.J., Forsberg, R., 1998. Reliability of F-scan in-
`shoe measurements of plantar pressure. Foot & Ankle Interna-
`tional 19, 668–673.
`Bartlett, R., 1999. Sports Biomechanics: reducing injury and improv-
`ing performance, E & FN Spon, London.
`Brandrup, J., Immergut, E.H., 1975. Polymer Handbook 2nd Edition.
`Wiley Interscience, New York.
`Cavanagh, P.R., Valiant, G.A., Misevich, K.W., 1984. Biological
`aspects of modelling shoe/foot interaction during running. In:
`
`MacNeil Exhibit 2179
`Yita v. MacNeil IP, IPR2020-01139, Page 7
`
`

`

`1386
`
`R. Verdejo, N.J. Mills / Journal of Biomechanics 37 (2004) 1379–1386
`
`ARTICLE IN PRESS
`
`Frederick, E.C. (Ed.), Sports Shoes and Playing Surfaces. Human
`Kinetics, Champaign, IL.
`Chen, H., Nigg, B.M., de Koning, J., 1994. Relationship between
`plantar pressure distribution under the foot and the insole comfort.
`Clinical Biomechanics 9, 335–341.
`Cook, S.D., Kester, M.A., Brunet, M.E., 1985. Shock absorption
`characteristics of running shoes. American Journal of Sports
`Medicine 13, 248–253.
`Dupont,
`1997.
`www.dupont.com/industrial-polymers/footwear/
`index.html. Thermal properties of Elvax measured by differential
`scanning calorimetry.
`Edington, C.J., Frederick, E.C., Cavanagh, P.R., 1990. Rearfoot
`motion in distance running. In: Cavanagh, P.R. (Ed.), Biomecha-
`nics of Distance Running. Human Kinetics, Champaign, IL,
`pp. 135–162.
`Gefen, A., Megido-Ravid, M., Itzchak, Y., 2001. In vivo behavior of
`the human heel pad during the stance phase of gait. Journal of
`Biomechanics 34, 1661–1665.
`Gross, T.S., Bunch, R.P., 1989. Discrete normal plantar stress variations
`with running speed. Journal of Biomechanics 22, 699–703.
`Henning, E.M., Milani, T.L., 1995. In-shoe pressure distribution for
`running in various
`types of
`footwear. Journal of Applied
`Biomechanics 11, 299–310.
`Henning, E.M., Milani, T.L., 2000. Pressure distribution measure-
`ments for evaluation of running shoe properties. Sportverletzung/
`Sportschaden 14, 90–97.
`Lemmon, D., Shiang, T.Y., Hashmi, A., et al., 1996. The effect of
`insoles in therapeutic footwear a finite element approach. Journal
`of Biomechanics 30, 615–620.
`Miller-Young, J.E., Duncan, N.A., Baroud, G., 2002. Material
`properties of
`the human calcaneal
`fat pad in compression:
`experiment and theory. Journal of Biomechanics 35, 1523–1531.
`
`Mills, N.J., Rodriguez-Perez, M.A., 2001. Modelling the gas-loss creep
`mechanism in EVA foam from running shoes. Cellular Polymers
`20, 79–100.
`Misevich, K.W., Cavenagh, K.W., 1984. Material aspects of modelling
`shoe/foot interaction. In: Frederick, E.C. (Ed.), Sports Shoes
`and Playing
`surfaces. Human Kinetics, Champaign,
`IL,
`pp. 47–73.
`Mueller, M.J., Strube, M.J., 1996. Generalizability of in-shoe peak
`pressure measures using F-scan system. Clinical Biomechanics 11,
`159–164.
`Nigg, B.M., De Boer, R.W., Fisher, V., 1995. A kinematic comparison
`of overground and treadmill running. Medicine and Scienc in Sport
`and Exercise 27, 99–105.
`Setybudhy, R.H., Ali, A., Hubbard, R.P., et al., 1997. Measuring and
`modelling of soft tissue and seat interaction. In: Progress with
`Human Factors in Automotive Design, Vol. SP-1242. Soc. Auto.
`Eng., Warrendale, PA, USA.
`Shorten, M.R., 2000. Running shoe design: protection and perfor-
`mance. In: Pedoe, T. (Ed.), Marathon Medicine. Royal Society of
`Medicine, London, pp. 159–169.
`Shiang, T.-Y., 1997. The non-linear finite element analysis and plantar
`pressure measurements for various shoe soles in heel region.
`Proceedings of the National Science Council, ROC, Part B: Life
`Sciences 21, 168–174.
`Taunton, J.E., Ryan, M.B., Clement, D.B., et al., 2003. A prospective
`study of running injuries. British Journal of Sports Medicine 37,
`239–244.
`Woodburn, J., Helliwell, P.S., 1996. Observations on F-scan in-shoe
`pressure measuring system. Clinical Biomechanics 11, 301–304.
`Verdejo, R., Mills, N.J., 2004. Simulating the effects of long distance
`running on shoe midsole foam. Polymer Testing, accepted for
`publication.
`
`MacNeil Exhibit 2179
`Yita v. MacNeil IP, IPR2020-01139, Page 8
`
`

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket