You are on page 1of 22

Mechanism and Machine Theory 140 (2019) 825–846

Contents lists available at ScienceDirect

Mechanism and Machine Theory


journal homepage: www.elsevier.com/locate/mechmachtheory

Research paper

Durability and design parameters of a Steel/PEEK gear pair


Damijan Zorko a,∗, Simon Kulovec b, Jože Duhovnik a, Jože Tavčar a
a
Faculty of Mechanical Engineering, University of Ljubljana, Aškerčeva 6, Ljubljana, Slovenia
b
Podkrižnik, d.o.o, Lesarska cesta 10, Nazarje, Slovenia

a r t i c l e i n f o a b s t r a c t

Article history: The parameters necessary for the reliable design of PEEK gears were determined. The ef-
Received 18 April 2019 fect of a steel gear’s surface on the polymer gear’s wear was investigated. Values for the
Revised 14 June 2019
wear coefficient of a steel/PEEK gear pair in dry and lubricated conditions are proposed.
Accepted 1 July 2019
Testing showed that, on average, grease lubrication extends the lifespan of PEEK gears by
Available online 5 July 2019
a factor of 1.23, while the additional treatment of steel gears with trovalisation and lubri-
Keywords: cation extends the lifespan by a factor of 2.54. Temperature measurements were used to
PEEK determine the coefficient of friction that is required to compute the operating temperature
Lifespan testing of a steel/PEEK gear pair. Using numerical simulations it was found that a proper tooth tip
Wear relief can improve the meshing and results in a further extension of the gear’s lifespan.
Surface © 2019 The Authors. Published by Elsevier Ltd.
This is an open access article under the CC BY-NC-ND license.
(http://creativecommons.org/licenses/by-nc-nd/4.0/)

1. Introduction

Polymer gears are used in ever-more-demanding operating conditions. This creates the need for high-performance poly-
mer materials, such as PEEK. This material is known for its excellent mechanical properties [1–3], which remain unchanged
at higher temperatures. The performance of PEEK has been widely studied by many researchers, with most of them perform-
ing standardised tribological [4–10] or mechanical tests [1,11,12]. Only a few PEEK gear applications were studied [13,14].
PEEK has also been studied as a material for rolling elements in bearings [15], for bearing rings [16] and in medicine, for
the artificial couplings in joint arthroplasty [5]. Greco et al. [4] evaluated the friction and wear behaviour of various PEEK
materials in high-speed contacts (up to 65 m/s) with stainless-steel balls using a 3-ball-on-flat contact configuration. Tests
were carried out on PEEK with no reinforcement, short-woven-fibre-reinforced PEEK and long-woven-fibre-reinforced PEEK.
Under the test conditions the material with long woven fibres exhibited the lowest wear rate and the lowest friction at
high sliding speeds. Schroeder et al. [6] analysed base PEEK, carbon-fibre-reinforced PEEK and carbon-fibre-reinforced PEEK
filled with graphite and PTFE using different tribological tests. The unfilled PEEK exhibited a very low scuffing resistance
and a large wear rate. Carbon-fibre-reinforced PEEK exhibited a very low scuffing resistance, but higher sliding and micro-
abrasive wear resistance. The addition of PTFE and graphite to the carbon-fibre-reinforced PEEK produced a sharp decrease
in the friction coefficient, high scuffing and an abrasion resistance with almost no wear. The increased performance ob-
served for the carbon-fibre-reinforced PEEK filled with graphite and PTFE was due to the transfer of a protective tribo-layer
from the composite to the counter-body and vice-versa. Mohammed and Fareed [7] studied the effect of a thin nanocom-
posite coating of ultra-high-molecular-weight polyethylene (UHMWPE) reinforced with carbon nanotubes. Ball-on-disc tests


Corresponding author.
E-mail address: damijan.zorko@lecad.fs.uni-lj.si (D. Zorko).

https://doi.org/10.1016/j.mechmachtheory.2019.07.001
0094-114X/© 2019 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license.
(http://creativecommons.org/licenses/by-nc-nd/4.0/)
826 D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846

Nomenclature

σF root stress ( mNm2 )


KF factor for tooth root load (-)
YFa form factor (-)
YSa stress correction factor (notch effect) (-)
Yɛ contact ratio factor for root stress (-)
Yβ helix angle factor for root stress (-)
Ft nominal tangential force (N)
b face width (mm)
bw common face width (mm)
mn normal module (mm)
Wm average linear wear (mm)
Td nominal torque (Nm)
NL number of load cycles (-)
HV degree of tooth loss (-)
kw wear coefficient (10−6 mm3 /(Nm ))
zi number of teeth (i=1 for pinion, i=2 for gear) (-)
lFl profile line length of the active tooth flank (mm)
u gear ratio (-)
βb helix angle at the base circle (βb = 0 for spur gears) (°)
ɛi partial contact ratio (i=1 for pinion, i=2 for gear) (-)
σH flank pressure ( mNm2 )
ZE elasticity factor (-)
ZH zone factor (-)
Zɛ contact ratio factor for flank pressure (-)
Zβ helix angle factor for flank pressure (-)
KH factor for tooth flank loading (-)
di reference circle diameter (i=1 for pinion, i=2 for gear) (mm)
Ca size of the tip relief (mm)
ϑFuβ root temperature (°C)
ϑ0 ambient temperature (°C)
P nominal output power (W)
μ coefficient of friction (-)
0,75
s K· ( m ) · mm1,75
kϑ, Fuβ heat-transfer coefficient of the polymer gear ( W )
vt tangential velocity (m/s)
R λ, G heat transfer resistance of the mechanism housing (K · m2 /W)
AG heat-dissipating surface of the mechanism housing (m2 )
ED relative tooth-engagement time with respect to ten minutes (-)
A initial point of tooth contact (-)
B lowest point of single tooth contact (LPSTC) for the drive gear and highest point of single tooth contact (HP-
STC) for the driven gear (-)
C pitch point (kinematic point) (-)
D highest point of single tooth contact (HPSTC) for the drive gear and lowest point of single tooth contact
(LPSTC) for the driven gear (-)
E end point of tooth contact (-)

with a standard stainless-steel ball were performed under dry conditions. The nanocomposite coating was found to be
very effective in improving the tribological properties of the PEEK. Rodriguez et al. [9,10] found that neat PEEK exhibits
a limited wear resistance against a metal counter surface. However, the wear resistance was significantly improved when
filling the PEEK with nano-sized particles of SiO2 . Regis et al. [5] investigated the applicability of PEEK and carbon-fibre-
reinforced PEEK in medical applications. The wear behaviour of three medical-grade PEEK formulations was tested under
dry and bovine-serum-lubricated conditions in a pin-on-flat test against Al2 O3 spheres. The crystallinity and the resulting
mechanical properties of the three selected materials were differentiated by annealing treatments. Reinforced formulations
exhibited an improved wear behaviour in dry and lubricated conditions. Shrestha et al. [11] researched the fatigue behaviour
of PEEK under uniaxial strain-controlled cyclic loading conditions, with various mean strains. Using a scanning electron mi-
croscope fractography analyses were performed on specimens under tensile mean-strain cyclic loading. Two types of inclu-
sions, unmelted particles and void-like defects, were identified to be the cause of the crack initiation. Void-like defects were
found to have a more detrimental effect on the fatigue life of the PEEK when compared to particles of a similar size. The
D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846 827

traditional design approach to reinforce PEEK is to use fibres for strengthening and filler particles for lubrication [6]. The
wear behaviour of the PEEK is enhanced by two major methods, composite preparation and surface modification [17]. The
frequency influence on the strain-controlled fatigue life of unfilled PEEK was found to be highly dependent on the strain
level. Increasing the load frequency at higher strain amplitudes resulted in longer fatigue lives [12].
To design polymer gears according to the available design methods in the VDI 2736 guidelines [18] the conversion pa-
rameters have to be determined by using the guidelines for the proposed gear testing procedure. Standard tribological and
mechanical tests provide a lot of information about how to increase material performance, but for a reliable gear design,
gear testing has to be performed. Tavčar et al. [19] found that fibre reinforcements do not always increase the gear’s lifespan,
although in simple mechanical testing the fibres do increase performance. A literature review yields the works of authors
who deal with different aspects of polymer gear testing; however, few of them deal with PEEK gears [13,14]. Mao et al.
[20,21] studied the wear of gears made of POM, and they also developed their own method for dimensioning POM gears.
Letzelter et al. [22] developed a dedicated gear-testing rig, where the sources of the heat in polymer gears can be pre-
cisely analysed. They tested PA66 gears. Kalin and Kupec [23] tested gears at controlled temperatures and compared the
results with those at uncontrolled temperatures. The results showed a significant influence of temperature on the lifespan
of POM gears. The contact conditions in meshing gears can be approximated by means of rolling and relatively-sliding cylin-
drical discs [24]. Wright and Kukureka [25] were investigating the appropriateness of simple pin-on-disk tribological tests
for predicting gear wear. They also attempted to simulate meshing gears by rolling and relatively sliding cylindrical discs.
Their results showed that these methods are not reliable for predicting gear wear. For reliable results, the best solution
is to make the gears from a particular material and test them under the operating conditions. As testing the lifespan of
polymer gears is time-consuming in practical terms, Pogačnik and Tavčar [19,26] developed a new, accelerated method for
testing polymer gears, which makes it possible to determine the critical load of a gear pair in a shorter time. Duhovnik et al.
[27] tested non-involute POM/PA gear pairs and compared their performance with the performance of involute gears. Mo-
han and Senthilvelan [28] investigated the bending strength of PP and PP+20%GF composite gears with asymmetric teeth.
Hasl et al. [29] tested oil-lubricated steel/POM gear pairs. They developed a method for calculating the tooth-root stress,
by taking into account the increase in the contact ratio due to load-induced teeth deflection [30]. When using lubrication
pitting, wear can occur. Berer et al. [15] investigated the pitting wear of injection-moulded PEEK rolls. The pitting occurred
only in the presence of lubricants. The authors performed a chemical analysis with Raman spectroscopy and found no effect
of the lubricant on the properties of PEEK, stating that PEEK is resistant to almost all common chemical influences. Using
a detailed microscopic analysis, they found that there were pre-cracks on the surface of the injection-moulded rolls. The
crack-growth mechanism was strongly supported by the fluid pressure of the lubricant enclosed in the pre-cracks. By adapt-
ing the injection-moulding process the surface quality of the PEEK rolls was improved and the fatigue lifetime of the rolls
was increased by a factor of 2–3.
Polymer gears fail due to different types of defects. The main types of failure include temperature-induced failure, wear
and fatigue [31]. The types of polymer-gear failures are varied and occur in typical operating conditions that depend on the
material pairing, lubrication and loads. For example, an identical material pair at a relatively high load will fail due to an
excessive temperature load, while at a lower load it will fail due to wear, and in the case of lubrication it might fail due to
fatigue, i.e., tooth fracture. It often happens that a combination of several types of failures occur, e.g., the material melts and
the tooth breaks or the tooth wears out and breaks at the root as a result of a reduced tooth thickness. Different authors
deal with different types of defects and the conditions in which they occur in different ways [26,32]. Testing experimental
gears provides an insight into what type of failure should be expected in a real-life application. Gears should be designed
according to the expected type of failure.
In the case of a steel and polymer gear combination, the softer surface is expected to wear out due to the great difference
between the hardness of the surfaces in the contact. There are three known wear mechanisms when a polymer is in contact
with a metal [33]. They include adhesive wear, abrasive wear and thermal wear. Intuitively it would be expected that the
lower roughness of a steel gear reduces the wear; however, studies have shown that the wear for extremely smooth surfaces
is comparable with that of relatively rough surfaces [34]. When a surface is too rough, it accelerates the polymer wear, as
the rough surface of a hard metal abrasively wears the polymer. Polymer wear depends on how deep the tips of the metal
surface penetrate, on the shear angle and on the sliding distance. In practice, the wear rate changes, as the valleys gradually
become filled with the removed polymer material forming a transfer layer, which reduces the penetration and consequently
slows the material erosion.
The surface topography of steel gears is dependent on the manufacturing technology. To improve surface properties ad-
ditional operations can be used, i.e. grinding, honing, and trovalisation [35,36]. Ramachandra and Ovaert [37] studied the
effect of controlled surface topography on the wear of Victrex PEEK 450 G in unlubricated conditions. They found that trans-
fer films play an important role in the steel/PEEK contact. Steel disks with different scribe configurations were tested. Pure
longitudinal, pure transverse, and yaw transversed configurations with different combinations of groove angle, spacing, and
height were produced. In the case of the surface roughness 0.2 μm rms, the groove angle was found not to have an effect
on the wear of PEEK, and the spacing effect could not be determined from the testing results. With increasing the groove
height, a larger effect on the wear rate was observed for the transversely scribed surfaces than for the longitudinally scribed
surfaces. Elliot et al. [38] studied the wear rate and friction of two PEEK grades, namely 450 G and 100P on a three-pin-on-
disk test. PEEK pins were run against stainless steel disks with two surface finishes, i.e. fine-grinding and polishing. Surface
roughness of the grounded disk was Ra=0.3 μm and for polished disk Ra=0.02 μm. PEEK grade 100P exhibited an almost
828 D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846

10 times larger wear against a ground surface than against a polished surface, but the coefficient of friction was larger
when running against a polished surface. In contrast, the wear of 450 G grade was almost independent of the two sur-
face finishes of the metal disk. Also, the coefficients of friction were more or less independent of the surface roughness.
Andrade et al. [39] studied the tribological contact between PEEK + 30%GF and metallic counterbody in oil lubricated con-
ditions. Metallic counterbodies were surface finished with four different finishing processes, i.e. turning, grinding, honing,
and polishing. A three-pin-on-disk test was used to research the correlation between the wear rate, coefficient of friction,
and lubrication regime. The wear rates measured for turned, grinded, and honed counterparts were very similar, while pol-
ished samples displayed much lower wear. This was due to different lubrication regimes, where polished samples operated
in a hydrodynamic lubrication regime, honed in mixed lubrication regime, while turned and grinded operated in boundary
lubrication regimes. No transfer film or debris was observed in any of the metal counterbodies, which was due to lubricant
removing the debris from contact region. Since grinding, honing, and polishing add a quite large additional cost in the pro-
duction part, an economically sustainable surface finishing process would be trovalisation (super finishing), where increased
efficiency has been recognised in the case of steel gear pairs [40]. When compared to grinding and honing the trovalisation
resulted in a smoother tooth surface without any asperity peaks and grinding lay [36].
The performance of PEEK in gear [13,14] and similar applications [41,42] can be enhanced with technological parameters
and by adding nanoparticles [43–45]. Zhang et al. [46] found that the wear resistance of PEEK can be enhanced by adding
graphitic carbon nitride. These studies [43–46] investigated the wear resistance of PEEK with tribological tests. The contact
conditions in a gear application are significantly different. In meshing gears, there is a combination of sliding and rolling
in the contact, and the normal force and the sliding speed at the point of contact are changing, while in the pin-on-disk
test, the normal force as well as the sliding speed are constant, and there is only sliding, with no rolling, in the contact.
Due to the low stiffness of polymer materials, the deflection of the polymer teeth in the mesh is significantly larger than
that of the metal gear teeth [30]. This also affects the conditions in the contact. The contact pressure, the sliding speed, the
amount of sliding and the temperatures affect the coefficient of friction of the steel/PEEK material pair [47,48], which makes
simple tribological tests inappropriate for predicting the lifespan of PEEK gears.
There are no available data for designing PEEK gears, neither in the latest VDI 2736 [18] guidelines nor in commercial
software. For this reason, the optimum and reliable design of PEEK gears requires data to be determined under test condi-
tions. In the presented research, steel/PEEK gear pairs were tested, where the steel gear was always used as the drive gear.
Tests were performed in both non-lubrication conditions and with the use of lubricating grease, in order to determine the
performance of the PEEK in a gear application when meshing with a steel gear. The objective was to obtain data on the
wear coefficient, the coefficient of friction and the fatigue strength, which would be used to design the PEEK gears.

2. Materials and methods

Due to the wide selection of different materials and the limited data for calculation, the reliable design and optimiza-
tion of polymer gears is a difficult task [49,50]. With this in mind, a design methodology was developed, consisting of
logical steps that lead towards the reliable use of polymer gears in a real-life application. Due to the high performance
requirements, the mass and noise standards, as well as the lubrication constraints, gear designers are often forced to use a
combination of polymer (also composite) and steel gears. It should be emphasised that developing new drives is always lim-
ited by the time and the costs, to which the design methodology, shown in Fig. 1, is adapted. Developing a new drive with
polymer gears often begins by using the materials for which the conversion data are available in the VDI 2736 guidelines
[18] or commercially available software, such as KISSsoft.
The VDI guidelines include strength-control data only for POM and PA66. It often happens that the customers’ require-
ments are too high, and gears made from these two basic engineering thermoplastics cannot fit into the required space. In
this case, a higher-performance material should be used. On the market, there is a huge selection of different polymer and
composite materials with different properties; however, there are mostly no corresponding data available for gear design.
Consequently, the calculation data must be obtained independently, based on time-consuming tests.
This paper presents research conducted during the development of a new polymer gear drive. The research began with
step 4 of the methodology presented in Fig. 1. The goal was to determine the design parameters for PEEK gears. When
designing a new steel/PEEK gear pair according the VDI 2736 [18] calculation methods the proper value of the coefficient
of friction is needed to perform the temperature calculation, the data on PEEK fatigue strength is needed to perform the
root strength control, and also the wear coefficient is needed to perform the wear control calculation. This additional testing
yields reliable gear design data, which saves time and the costs of developing a drive. Together with the tests, numerical
simulations were also conducted. A numerical model was developed, and it can be used to accurately calculate the stress in
a gear at all the meshing points, as well as to analyse the effects of geometric changes on the lifespan of gears.

2.1. Data on tested gears

Polymer gears in high-volume and mass production are manufactured by injection moulding, where the material’s shrink-
age during cooling should be taken into account [51]. For the purpose of better accuracy and due to the small number of
pieces, the gears tested in this research were hobbed. The steel gears were made from 42CrMo4 material and plasma ni-
trided. The polymer gears were made from Victrex PEEK 650 G material, the properties of which are shown in Table 1 [52].
D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846 829

Fig. 1. Polymer-gear-drive development methodology.

Table 1
Material properties of Victrex PEEK 650 G [52].

parameter value

tensile modulus (at 23 °C) 3900 N/mm2


tensile strength (at 23 °C) 95 N/mm2
flexural strength (at 23 °C) 155 N/mm2
melting-point temperature 343 °C
glass-transition temperature 143 °C
density 1.30 g/cm3
830 D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846

Fig. 2. Flank surface after: a) hobbing, b) hobbing and trovalisation.

The PEEK gears were hobbed from the base of a sliced extruded bar. Two different steel gears were used for the testing
(Fig. 2):

a) a hobbed gear with no additional mechanical treatment of the flanks, roughness: Ra = 0.761 μm
b) a hobbed gear, additionally treated with trovalisation (trovalisation time 1 h), roughness: Ra = 0.689 μm

Trovalisation is a surface treatment in a vibratory machine using abrasive material. The purpose of trovalisation is to
remove the sharp edges that remain after cutting. The steel gear’s flank profiles were measured using the MarSurf XC20
contourograph. Measurements of the flank geometry showed that the flank profile does not change during trovalisation.
Roughness measurements showed a minor effect of trovalisation on the surface roughness. The flank surface roughness of
steel gears was measured on a Tesa Rugosurf 90 G measuring machine, with the measuring direction being along the flank
profile. Three measurements of roughness were performed on each gear, and the mean value is used in this article. The
surface of the steel gears was analysed using a Keyence VHX-200 digital microscope. This microscope was also used to take
the photographs that were used for measuring the wear of the PEEK gears.
The geometric parameters of the test gears are collected in Table 2. According to DIN 3967 [53], the prescribed tooth
thickness allowance was e25, with a tooth-thickness deviation of −0.030/ −0.060 mm. The deviations of the finished gears
were measured with a Wenzel LH 54 3D measuring machine. Following the results, the steel and PEEK gears were classified
under quality grade 10, according to ISO 1328 [54]. The geometry of the test gears was defined according to the test rig’s

Table 2
Geometric parameters of the test gears.

parameter value

profile involute, ISO 53 A


module - m [mm] 1
number of teeth - z 20
pressure angle - α [°] 20
facewidth - b [mm] 6
coeff. of profile shift - x1 0
coeff. of profile shift - x2 0
transverse contact ratio - ε α 1.557
D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846 831

Fig. 3. Test rig and area of the temperature measurements, presented gear pair was tested at 1.2 Nm and 1600 rpm and failed after 5.94 · 106 cycles.

restrictions. The test rig was designed to allow the testing of larger gears; however, the maximum test load is 1.75 Nm.
Performing tests at this load until the breaking point on larger gears would take too long.

2.2. Test conditions

All the tests were performed in a laboratory environment at a temperature of 23 °C ± 3 °C and a room humidity of 45%
± 5%. The objective was to examine how the lubricated flank surface of a steel gear affects the lifespan of PEEK gears. To
do so, three different combinations were tested:
1. steel gear/PEEK gear, no lubrication,
2. steel gear/PEEK gear, lubricated with ISOFLEX TOPAS NB 52 grease,
3. trovalised steel gear/PEEK gear, lubricated with ISOFLEX TOPAS NB 52 grease.
In all of the tested examples, the steel gear was the drive gear and the polymer gear was driven. This is the most typical
example used for reducing the rotational speed in applications. Non-lubricated gear pairs were tested at three different load
levels, i.e., 0.8 Nm, 1.0 Nm and 1.2 Nm, and the lubricated gears only at 1.0 Nm and 1.2 Nm, as they achieved long lifespans
already under these loads. Tests were conducted at the rotational speed of 1600 rpm. Prior to operation, the lubricated
gear pairs were lubricated with ISOFLEX TOPAS NB 52 grease. Both the steel and the polymer gears were lubricated. No
fresh lubricant was fed during the test. All the tests continued until the PEEK gear failed. The tests were performed on
an in-house-developed test rig, Fig. 3. It is designed to allow free access to the gears to measure the temperature with a
thermal camera. The test rig allows a continuous torque setting, via a drive motor and a brake motor, controlled by frequency
regulators. The drive and the brake motors were powered by Siemens four-pole asynchronous electric motors with a nominal
torque of 2.5 Nm. The power is transmitted from the drive electric motor to the gear driveshaft via a toothed belt, which
has a beneficial effect on dampening the electric motor’s vibrations. Tests required the centre distance between the tested
gears to be precisely adjustable. For this purpose, a positioning stage was used, allowing adjustments to the centre distance
to within 0.02 mm. Before the test, once a gear had been mounted on the shaft, its run-out was measured. The acceptable
run-out was 0.05 mm and all the test gears were within this value. The run-out measuring tool’s precision was 0.01 mm.
Using an FLIR T 420 thermal camera, the gear operating temperature was measured for each load level, which showed the
effect of the grease on heat generation and consequently the gear temperature. The highest temperature over an area of
3 × 3 pixels was measured, located in the PEEK gear’s root diameter area, Fig. 3. For the temperature measurements and
once the thermal camera had been calibrated, the emissivity was set to ε = 0.95.

2.3. Numerical model

In order to obtain a view from another perspective, all the tested examples were also simulated with the finite-element
method. A numerical simulation was used to calculate the root stress and the flank pressure occurring in gears under the
test loads. The actual transverse contact ratios were determined, and they are larger than the theoretical values, which is
caused by a severe deflection of the polymer gear’s teeth [55]. Simulations were carried out in the Ansys Workbench 17.2
software. The geometrical model had 5 teeth, and the entire analysis of the results was performed on the middle tooth, as
it passes all the characteristic meshing points, Fig. 4. When modelling the gear geometry, the tooth-thickness tolerance was
taken into account, and the gear geometry was modelled at the centre of the tolerance area.
The material properties of both materials were modelled as linear elastic, taking account of the PEEK material’s param-
eters: tensile modulus EPE E K = 3900 N/mm2 and Poisson’s ratio νPE E K = 0.385. The friction between the tooth flanks was
also considered, and μ = 0.2 was used as a non-lubricated example and μ = 0.09 [18] as an example for greased lubri-
cation. The simulations were run in 2D, in the planar-stress state. PLANE183 finite elements were used with a quadratic
832 D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846

Fig. 4. Numerical model.

Fig. 5. Mesh refinement in the region of interest.

approximation of the displacement. The contact conditions were simulated with CONTA172 and TARGE169 elements. In the
contact areas between the tooth flanks and in the tooth root, the mesh was refined in order to achieve the highest possible
accuracy for the stress calculation, Fig. 5. The average quality of the finite-element mesh was 0.96. When the friction is
modelled and large deflections are taken into account, the simulations become non-linear, which significantly increases the
problem-solving time. The solution time for the presented numerical model is about 45 min.

3. Results

3.1. Lifespan

Fig. 6 shows the lifespan curves for the tested gear pairs. The curves are drawn for a 90% survivability of the test items.
This limit is often referred to as B10. The 90% survival rate was determined by means of the Weibull distribution, which was
used because it is suitable for a small number of samples [56]. To determine the β and η Weibull shape parameters, the
MINITAB 18.0 software was used. Ten percent of the working gear pairs at a torque of 1.2 Nm with no added lubricant will
fail after 4.62 · 106 cycles. At a lower torque of 1.0 Nm, the limit for a 90% survivability of a gear pair is 8.22 · 106 cycles. The
lowest testing load for non-lubricated gear pairs was 0.8 Nm. Under this load, 10% of the gear pairs will fail after 11.86 · 106
cycles or earlier. The PEEK gears achieved a longer lifespan, when compared to the POM and PA gears tested under identical
loads [19,32]. It was found that a steel/PEEK gear pair reaches significantly more cycles to failure even without lubrication.
The objective to extend the lifespan resulted in the test examples shown in Section 2.2. In the case of the added grease
lubrication and under a torque of 1.2 Nm, the B10 lifespan limit was 5.43 · 106 cycles, and at 1.0 Nm it was 10.42 · 106 cycles.
D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846 833

Fig. 6. Lifespan curves for the tested gear pairs (curves are drawn for a 90% survival rate).

Table 3
Extending the lifespan, compared to hobbed gears with no additional flank treatment and no added lubricant.

Torque [Nm] Nom. tang. force Ft [N] Root stress σ F [N/mm2 ] Lifespan-extension factor Mean lifespan-extension factor

Lubricated

1.2 120 66.56 1.18 1.23


1 100 55.46 1.27
Lubricated + trovalised

1.2 120 66.56 2.30 2.54


1 100 55.46 2.78

An additional treatment of the steel gears with trovalisation and the added lubricant further extended the lifespan. For these
gear pairs, the B10 limit is at 10.63 · 106 cycles for a torque of 1.2 Nm, and 22.80 · 106 cycles for 1.0 Nm.
Table 3 shows relative comparisons of the lifespan extension for PEEK gears in a mesh with a steel gear, whose surface
has been additionally treated and lubricated. It was found that applying a grease lubricant extends the PEEK gears’ lifespan
by a factor of 1.23, i.e., 23%. Compared to hobbed gears, working without lubrication, the additional treatment of a steel gear
with trovalisation plus lubrication extend the PEEK gears’ lifespan by a factor of 2.54. This shows that the lifespan extension
is greater at lower loads. At higher loads, the heat generation is higher, the gears work at a higher temperature, which then
increases their wear rate. The test results show that treating steel gears with trovalisation provides a major contribution
towards extending the lifespan of a steel/PEEK gear pair.
The root stress values of the tested gear pair were calculated in the KISSsoft 03/2015 software, and this equation, recom-
mended by the VDI 2736 guidelines, was used:
Ft
σF = KF · YF a · YSa · Yε · Yβ · (1)
b · mn
The following correction coefficients were used for the tested PEEK gears’ geometry: KF = 1.0, YF a = 3.01, YSa = 1.51,
Yε = 0.732, Yβ = 1.

3.2. Temperatures during operation

The gears’ surface temperature was measured in all the tests. The polymer-gear temperatures, measured throughout the
test, are shown in Fig. 7. The measured temperatures were used to determine the coefficient of friction at the teeth contact
in the steel/PEEK gear pair, which is explained later in the discussion. When the lubricant was used, the temperature of
the polymer gears was approximately 20 °C lower, compared to the non-lubricated tests, which is due to a lower coefficient
of friction and better heat dissipation via the lubricant. In the lubricated case, the additional treatment of the steel gear
with trovalisation had no significant effect on the operating temperature of the gear pair. The temperatures were essentially
identical at the same torque levels, irrespective of how the steel gear had been treated, Fig. 8. In this case, the only effect
on the temperature was the torque transfer through the gear pair. Compared to the material combination polymer/polymer
[32], the measured surface temperatures are significantly lower at identical loads. Lower temperatures at identical loads
are due to the higher thermal conductivity of the steel, compared to the polymer, which significantly improves the heat
dissipation.
The temperature increases during all the tests, which means that heat generation surpasses its dissipation. This results
in heat accumulation and an increasing gear temperature. In the 1.0 Nm tests with added lubrication, the temperature
remained steady for the major part of the test, but then begins to increase. It can be concluded that the increase is the
834 D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846

Fig. 7. Operating temperature of test gear pairs, steel gear not additionally treated.

Fig. 8. Temperature of lubricated gear pairs.

result of gear wear. The meshing conditions in the worn gears are no longer the same as they were at the beginning of
the test. In the worn gears, the heat generation increases, which can result from several factors, such as: changes in the
coefficient of friction, changed force in the contact and changed sliding conditions in the contact.

3.3. Types of failure

Wear was the main reason for failure in all of the tested gears. In the lubricated as well as the non-lubricated gears,
the flanks of the polymer gears were heavily worn, Fig. 9. Wear advances to the point of critical tooth thickness, and when
this is reached, the tooth breaks. In several tests, cracks in the dedendum flank area were also spotted; this occurs due

Fig. 9. PEEK-gear wear during operation with no added lubrication (torque 1.0 Nm, rotational speed 1600 rpm, wear after 9.13 · 106 cycles).
D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846 835

Table 4
The wear coefficient values [10−6 mm3 /(Nm )].

material pair kw

steel-PEEK (Ra=0.761 μm), dry 6.84


steel-PEEK (Ra=0.761 μm), lubricated 5.43
steel-PEEK (Ra=0.689 μm), lubricated 3.77

to the reduced tooth thickness. Wear has a major effect on the noise and vibrations, caused by the gear pair [57]. In the
increased wear conditions, noise and vibrations exceed what is acceptable for use in real-life applications. For this reason,
close attention should be paid to the wear control in polymer/steel gear combinations, both lubricated and non-lubricated.

3.3.1. Wear control


The VDI 2736 guidelines suggest a wear-control model for gear pairs. The following equation is used to calculate the
mean linear wear:
Td · 2 · π · NL · HV · kw
Wm = ≤ 0.2 · mn (2)
bw · z · lF l
In VDI 2736 the wear coefficients kw are specified only for the POM and PBT materials for two roughness values of the
paired steel gear. For designing the PEEK gears, test results were used to determine the wear coefficient for PEEK in a mesh
with a steel gear. Using Eq. (2), the wear coefficient is expressed as follows:
Wm · bw · z · lF l
kw = (3)
Td · 2 · π · NL · HV
For known values of the parameters in Eq. (3) and the lifespan test results:
Td = value is in Table 7, bw = 6 mm,
NL = lifespan test results, Fig. 6 z = 20,
HV = 0.206, lF l = 1.673 mm.
The calculated wear coefficients are collected in Table 4. The method of determining the wear rate Wm [mm] is shown
in Fig. 10. The degree of tooth loss HV is a factor included in the VDI 2736 [18] wear and temperature calculation model,
which describes the influence of the tooth geometry on the frictional losses and consequent heat generation. Different tooth
geometries result in different contact ratios. The greater is the contact ratio the longer the teeth are in contact and, hence,
the greater are the frictional losses. The degree of tooth loss is calculated using the following equation:
π · (u + 1 )  
HV = · 1 − ε1 − ε2 + ε12 + ε22 (4)
z2 · cos βb
The worn PEEK gears were photographed at 50-times magnification and imported into the Siemens NX 12.0 modelling
software. There, the worn gears were compared to the theoretical geometry of gears, on the basis of which the hobbing
tools were manufactured. The longest distance, perpendicular to the theoretical flank and running between the worn flank
and the theoretical shape of the flank profile, was measured. The wear was measured on several teeth of the same gear,
and from the results the mean value Wm was calculated for each gear. For this mean value and using Eq. (3), the wear
coefficient for each gear was then calculated. For the calculated wear coefficients of each gear, the mean wear coefficient

Fig. 10. Determining the wear rate Wm .


836 D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846

Fig. 11. Root stress (left) and flank pressure (right) for the driven PEEK gear, calculated for non-lubricated test examples.

was then determined for all of the gears, tested on the test rig under identical lubrication conditions and roughness of the
steel gear’s surface.

3.4. Numerical results

All the test examples were simulated using a numerical model with a frictional contact between the meshing flanks.
The main point of interest was the stress in the PEEK gears during meshing. Fig. 11 shows the course of the highest root
stress and flank pressure, calculated for the non-lubricated examples (μ = 0.2 ). The calculated root stress and flank pressure,
based on the numerical model, differ from the theoretical course of the stress [35]. The theoretical models for root stress
calculation in VDI2736 [18] (for polymer gears) and ISO 6336 [58] or DIN 3990 [59] (for steel gears) take into account
only the tangential force acting on the tooth. Friction between the meshing flanks is neglected in those models. Also, the
geometry of the gears is considered as rigid, not taking into account the teeth deflection. According these models, the
maximum root stress in the driven gear ocurrs in point B, which is the HPSTC for the driven gear. The lever arm from the
meshing point B to the critical cross section in the tooth root is the largest in that case, leading to maximum root stress.
In the case of frictional contact, two forces are acting on the tooth, i.e. the tangential force and the frictional force. The
direction of the frictional force acting on the tooth of the driven gear is always towards the pitch point C [20,25]. In point B
the horizontal component of the frictional force and the tangential force acting on the tooth are acting in opposite directions,
resulting in a lower root stress than in a case of a frictionless model [60]. The highest simulation calculated root stress in
the driven PEEK gear therefore occurs at point D, where the tangential force on the tooth Ft and the horizontal component
of the frictional force between the meshing teeth are acting in the same direction. Considering gear as a deformable body
leads to a decrease in the single tooth contact area, due to teeth deflections, especially in the case of polymer gears. With
the reduction of the single tooth contact, the lever arm, from the point B to the critical cross section in the root, is smaller.
As the analytical models [18,58,59] do not consider these two effects, the highest root stress in the driven gear occurs at
point B.
In the flank pressure, peaks can be observed at the beginning and the end of the meshing, at points A and E. Flank
pressure also increases in the area of a single-pair tooth contact, between points B and D, where the entire load is trans-
ferred through one pair of teeth. Fig. 12 shows the root stress and the flank pressure in the lubricated case, where a lower

Fig. 12. Root stress (left) and flank pressure (right) for the driven PEEK gear, calculated for lubricated test examples.
D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846 837

Table 5
Calculated root stresses and flank pressures.

moment root stress root stress flank pressure flank pressure


load (VDI 2736) (simulation) (VDI 2736) (simulation)
[Nm] [N/mm2 ] [N/mm2 ] [N/mm2 ] [N/mm2 ]

0.8 44.37 42.97 97.42 95.32


1.0 55.46 51.13 108.92 107.40
1.2 66.56 58.27 119.31 118.86

coefficient of friction (μ = 0.09 ) between the meshing flanks was used. Flank pressure is slightly lower in the case of lubri-
cation; however, there are still pronounced peaks at the beginning and the end of the meshing. These pressure peaks are
not part of a standard calculation [18]; however, numerical simulations show that they are present. The standard calculation
considers gears as non-deformable bodies, so no teeth deflection is taken into account. In reality, due to teeth deflection,
the path of contact is not a straight line anymore, at the start of the meshing there is a premature contact where the tip of
the driven gear starts meshing with the dedendum part of the drive gear [41]. The radius of the tip is considerably smaller
than the flank curvature in other meshing points below the tip, leading to a larger contact pressure. At end of the meshing,
the contact is prolonged and this time the tip of the drive gear is in contact with the dedendum part of the driven gear.
The tip rounding radius of the driven gear is also considerably smaller than the flank curvature of other meshing points on
the flank below the tip, leading to increase in flank pressure. In the flank area, where the pressure peaks occur, the highest
sliding speeds also occur. Wear depends on both the contact pressure and the sliding speed, and if both maximums occur
in the same area, the highest wear is to be expected there, which correlates with the test results.
For a back-to-back comparison with the results of the numerical calculation, the root stress and flank pressure were also
calculated in the KISSsoft 03/2015 software, Table 5. The calculation followed the VDI 2736 guidelines model. To calculate the
root stress, Eq. (1) was used with the values of the correction coefficients, explained in Section 3.1. The following equation
was used to calculate the flank pressure:

Ft · KH u + 1
σH = Z E · Z H · Z ε · Z β · · (5)
bw · d1 u
For the tested gear pair, the values of the equation coefficients are as follows: ZE = 34.476, ZH = 2.495, Zε = 0.902, Zβ = 1,
KH = 1, bw = 6 mm, d1 = 20 mm, u = 1.
The simulated-calculated stress values were compared with the VDI 2736 guidelines calculation. For the root stresses,
the highest calculated root stress that occurs at point D was compared. For the flank pressures, it was the highest calculated
value in the area of a single-pair tooth contact, i.e., between points B and D. The VDI 2736-calculated stresses correspond
well with the numerical simulated-calculated stresses. The correspondence is the closest under a load of 0.8 Nm. With in-
creasing loads, the differences between the simulated and the VDI 2736-calculated stresses are increasing too. The conducted
simulations considered the increased transverse contact ratio as a result of tooth deflection and the friction between mesh-
ing flanks, while the guidelines take no account of these two effects. For this reason, the numerical simulated-calculated
root stress values are slightly lower [30,60].
For the stress calculations, the presented numerical model, as well as the VDI 2736 model, take no account of the effect
of temperature. During operation, the temperature of a gear pair increases, which reduces the stiffness of the polymer
material. Reducing the stiffness further increases the transverse contact ratio. Since the effect of contact ratio on the gears’
stress state is taken into account with contact ratio factors for the root stress (Yɛ ) and flank pressure (Zɛ ), its increase
should reduce the values of the contact ratio factors, Yɛ in Eq. (1) and Zɛ in Eq. (5). A lower stiffness of the parts in the
contact should also have some effect on reducing the value of the elasticity factor ZE in Eq. (5). Factor ZE takes into account
the modulus of elasticity, and Poisson’s ratio of the gears in contact when calculating the flank pressure. With the rise in
temperature, the modulus of elasticity is reduced and consequently also the value of the factor ZE should be reduced, leading
to a lower flank pressure. Not considering these effects makes the VDI 2736 and numerical model-based stress calculations
conservative. When designing polymer gears one should be careful, because with the increase of the temperature also the
fatigue strength of the polymer material is reduced.

4. Discussion

4.1. Restrictions on the use of the results

The results were obtained from testing the gears’ test geometry, under the above-explained test conditions. The geometry
of the gears in a realistic drive application will, in most cases, be different from the geometry of the test gears (different
module, number of teeth, tip relief, pressure angle). When using the presented results for dimensioning PEEK gears, we
should be careful and take account of the factors that are explained below in more detail. In any case, when a new drive
is being developed, several prototypes should be made and a realistic size of the gears under realistic operating conditions
should be tested. If the gears are not properly dimensioned from the outset and the prototype testing does not yield the
838 D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846

required lifespan, a lot more work and costs will follow later, as altering gear dimensions will also cause changes in other
parts of the drive [61]. Changed gears also require a new injection-moulding tool. The more a drive is sophisticated, the more
parts need to be changed. In the best-case scenario it will mean extra work, but most often it will incur extra manufacturing
costs. By testing experimental gears and determining the parameters for designing them with existing methods [18], it is
possible to properly design gears as early as the concept stage, which saves time and reduces the drive’s development costs.

4.1.1. Effects on fatigue strength


The fatigue strength of a PEEK gear is a function of the material, dependent on the temperature and the frequency
of the load. It is not directly dependent on the geometry of the gears, the treatment of the steel gear in the contact, or
the lubrication. At a certain stress in the material, at a certain temperature and at a certain frequency of the load, a gear
will survive a certain number of cycles. This can lead to the conclusion that the acquired data on fatigue strength also
apply to other gear geometries. Gear lifespans, shown in Fig. 6, depend on both the treatment of the steel gear and the
type of lubrication. This is due to the type of gear failure, which in all of the tests was caused by wear, and the different
temperatures at which the tested gear pairs worked. Consequently, our tests cannot be used to directly determine the fatigue
strength of PEEK, as wear and not fatigue was the type of failure. Teeth broke when the gear failed due to the reduced tooth
thickness, which caused increased stress in the dedendum flank area. PEEK gears are expected to exhibit a longer lifespan
in an application where there is no wear. If the results shown in Fig. 6 are to be used to determine PEEK’s fatigue strength
it will mean a conservative approach. The effect of temperature on PEEK’s fatigue strength should also be considered. In the
conducted tests, the gears operated at different temperatures. Dry gear pairs worked in the temperature range between 55 °C
and 65 °C, and the lubricated ones in the range between 35 °C and 45 °C. The temperature of a particular gear pair mostly
depended on the torque transferring through it. Applying a conservative approach, a specific fatigue strength from the lower
part of the temperature range should be used for gear design. This means that the PEEK’s fatigue strength obtained when
testing dry gear pairs applies to a temperature of 55 °C and for the lubricated gears to 35 °C. For more accurate values, one
should look at what temperature and under which load a test yielded a certain number of cycles. Gears are designed on
the safe side when a gear pair in a drive that is being developed is operating at a temperature lower than the temperature
for which the fatigue strength was determined. Tests were carried out at the rotational speed reported in Section 2. As the
frequency of the load affects the PEEK’s fatigue strength [12], the results apply to the frequencies of the load within the
range of the performed tests.

4.1.2. Effects on the wear coefficient


The wear coefficient was determined from the wear-control equation suggested in VDI 2736. The equation considers
the effect of the number of teeth and the face widths, as well as the active lengths of the tooth flank. It directly depends
on the module. The coefficient HV also depends on the number of teeth and the module. Such determined values of the
wear coefficient should also apply to other gear geometries. The wear control Eq. (2) does not directly take account of the
effect of temperature on the mechanical properties of a polymer material. The effect of the temperature on the coefficient
of friction is covered by the wear coefficient; however, it is not specified in the VDI 2736 guidelines to what temperatures
the suggested values of the wear coefficients apply. The value of the wear coefficient determined in this research applies
to a temperature of 55 °C for dry gear pairs and 35 °C for lubricated gear pairs. The coefficient of friction for the steel/PEEK
combination depends on the contact pressure, the sliding speed, the sliding length and the temperature. When the gear
geometry is significantly larger than the tested pair, a larger module and more teeth affect the temperature during gear
operation, and the coefficient of friction as well as the wear coefficient will change as a result. The presented parameters
serve the purpose of dimensioning the gears for operation under conditions similar to the tested values. For the optimum
geometry and reliable functioning, it is vital to always test gears of a size similar to the one that will be used in the drive,
and under the conditions that are similar to the drive’s operating conditions.

4.1.3. Effect of manufacturing technologies


It should be noted that the performance of polymer gears also depends on the manufacturing technology used to man-
ufacture them. The test PEEK gears were manufactured from extruded bars by means of cutting. Transferring the specified
fatigue strength, wear coefficient and coefficient of friction onto the injection-moulded gears should therefore include some
reservations. The injection moulding of PEEK is a demanding job, requiring extensive experience of the injection-moulding
process itself (injection pressure, temperature, molten material flow rate) as well as of tool making (gate positioning). If the
PEEK injection moulding is poorly designed, the performance of poorly manufactured gears [11,15] will not match the gears
presented in this paper.

4.1.4. Effect of gear-geometry deviations


According to the measured deviations, the tested gears were classified as ISO 1328 quality grade 10. The effect of devia-
tions on the root stresses in metal gears is considered in the coefficients KFα and KFβ [58]. The effect of deviations on the
lifespan of polymer gears in a mesh with steel gears has not yet been systematically researched. VDI 2736 suggests that due
to the elastic properties of thermoplastics and when b/mn < 12, the coefficients KFα and KFβ should not be considered, i.e.,
their value equals 1. Load transfer through the gears is also affected by the assembly deviations [62]. It can be assumed that
loads between the gears in the contact between a steel and a polymer gear will spread more evenly when a more accurate
D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846 839

geometry has been achieved. Due to its large tensile modulus, a steel gear is not as flexible as a polymer one. Therefore, an
inferior accuracy can lead to a concentrated stress, while higher accuracy spreads the load more evenly between the teeth.

4.2. Effect of trovalisation

Together with a lubricant, the additional treatment of steel gears by trovalisation also reduces the wear. It did not sig-
nificantly change the flank surface’s roughness; however, it removed any sharp edges and improved the roundness of the
tooth tips on the steel tooth. Using the digital Keyence VHX 200 microscope, photographs were taken at 200-times magnifi-
cation, where differences between the surfaces of steel gears can be observed. Fig. 13 shows a steel gear with no additional
treatment after hobbing, and Fig. 14 shows a steel gear that was trovalised after hobbing. It is clear that the sharp edges
were removed on the trovalised gear, and the tip itself has a larger fillet radius. The tip fillet radius was assessed in the
modelling software, where the measured radius at the tip with no additional treatment was R=0.03 mm, while at the tip of
the trovalised gear it was R=0.3 mm. The effect of tooth-tip rounding was analysed by means of numerical simulations. It
was found that increasing the tip rounding significantly lowers the flank’s pressure peaks, which is shown in Fig. 15. When
the fillet radius increases from R=0.03 mm to R=0.3 mm, it reduces the flank pressure at the beginning of the mesh by 48%
and by 38% at the end. Sliding speeds are also at their highest at these points, which means that a reduced contact pressure
significantly reduces the wear. At the beginning of the mesh, the tip of the driven polymer gear is in contact with the root
of the drive gear’s tooth. It is expected that the tips of the polymer gears will gradually wear out, which will reduce the
contact pressures. At the end of the meshing, the tip of the steel gear is in contact with the root of the polymer gear. As the
harder steel gear is not expected to wear, a larger tip-fillet radius should be created when making the steel gear, and sharp
edges removed by an appropriate additional treatment – trovalisation. Rounding the tooth tip has no significant effect on the
root stresses in the gear, Fig. 15. Increasing the fillet radius can significantly reduce the wear rate and increase the lifespan.
This conclusion correlates with the experimental results, as the PEEK gears meshing with trovalised steel gears exhibited
a much longer lifespan. In the performed test cases rounding the tip had a more significant effect on the wear than the
surface roughness. For hobbed gears, the measured roughness equalled Ra=0.761 μm, and for hobbed and then trovalised
gears it was Ra=0.689 μm. There is a difference in the roughness, but it is small. In this case the effects of roughness and
tip rounding cannot be accurately distinguished.

Fig. 13. Steel gear, no additional treatment, front view (top), axonometric view (bottom), 200-times magnification.
840 D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846

Fig. 14. Finished steel gear, trovalised, front view (top), axonometric view (bottom), 200-times magnification.

Fig. 15. Effect of tip-fillet radius on the root stress and flank pressure of the driven PEEK gear (black dashed lines represent meshing points for the gears
with tip rounding R=0.03 mm, red dash-dotted lines represent meshing points for the gears with tip rounding R=0.3 mm).

4.3. Measures to reduce wear

The measures to reduce wear include lubrication, proper gear-surface roughness (especially for steel gears) and tooth ge-
ometry that minimises the specific sliding [63]. Tooth-tip rounding also showed a significant effect. Numerical simulations
showed that the size of the tip-fillet radius affects the flank pressure occurring in the areas of the highest sliding speeds.
Simulation results also correlate well with the test results, where the gears with a larger tip rounding achieved longer lifes-
pans. In practice, different methods are used to improve the load transfer [64,65]. Most of them were developed for metal
gears. A familiar measure is adjusting the flank profile where either the tip or the root of the tooth can be adjusted. The
lead profile is also often adjusted in order to obtain a convex shape of the flanks across the gear’s width. Using numerical
D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846 841

Fig. 16. Analysed tip reliefs.

Fig. 17. Effect of tip relief on the root stress and flank pressure of the driven PEEK gear, torque 0.8 Nm (red dashed lines represent meshing points for
the gears with no tip relief, green long dash-dotted lines represent meshing points for the gears with tip relief Ca =0.015 mm, blue short dash-dotted lines
represent meshing points for the gears with tip relief Ca =0.030 mm).

simulations, the effect of tip relief on the flank pressure in the gear was tested. Two different tooth-tip modifications were
used. The tip reliefs in the sizes of Ca =0.015 mm and Ca =0.030 mm were analysed, extending out to the diameter of the
highest point of a single-pair tooth contact, Fig. 16. In all of the simulated examples, the tip relief was modelled on both
the drive and the driven gear.
Fig. 17 shows a comparison between the root stress and the flank pressure under a load of 0.8 Nm. It transpires that
a properly selected tip relief can significantly improve the flank-pressure conditions at the beginning and the end of the
meshing. The selected reliefs have no meaningful effect on the maximum root-stress value, only that its course changes
slightly. A numerical simulation was used to calculate the tooth deflection, which made it possible to determine the real
transverse contact ratios. From the simulation results, the characteristic meshing points A, B, C, D, and E for the deflected
teeth geometry were determined. The actual contact ratio was calculated as the quotient of the meshing angle between
points A-E, and the meshing angle between points A-D. The maximum tooth deflection on the tip diameter in the gear’s
circumferential direction was analysed, Table 6. At a torque of 0.8 Nm, the calculated tooth deflection was 0.043 mm. With
a tip relief of Ca =0.030 mm the flank-pressure peaks no longer occur, which shows that the tip relief should be of a size

Table 6
Calculated real transverse contact ratios.

torque no tip relief Ca =0.015 Ca =0.030 deflection

0.8 Nm 1.792 1.723 1.573 0.043 mm


1.0 Nm 1.823 1.762 1.672 0.052 mm
1.2 Nm 1.846 1.792 1.710 0.059 mm
842 D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846

Table 7
Power transmitted through the gear pairs in
our test examples.

P [W] Td [Nm] n [rpm] v [m/s]

197.54 1.2 1572 1.65


167.76 1.0 1602 1.68
136.47 0.8 1629 1.71

that is similar to the tooth deflection. The theoretical transverse contact ratio of the tested gear pairs is εα =1.557. For the
tested examples, the calculated real transverse contact ratios are shown in Table 6. It is clear that the tip relief marginally
reduces the transverse contact ratio. However, in all the simulated cases, the real transverse contact ratios were larger than
the theoretical value, that applies to a non-deformable geometry. A larger transverse contact ratio has a beneficial effect
on the stress conditions in the gear, as the area of a single-pair tooth contact is narrower. The negative aspect of a larger
transverse contact ratio is a longer duration of the mesh, which increases the losses due to friction in the teeth contact.

4.4. Temperature calculation

Operating temperature is a vital factor that requires attention when designing a gear pair. When a polymer gear heats
up beyond the acceptable temperature for continuous operation it leads to thermal failure of the gear. There were several
numerical [66–68] and analytical [20,69] methods proposed for the polymer gear temperature calculation, but the vast ma-
jority of them is rather complex to be used in practice by gear design engineers. The VDI 2736 guidelines [18] on the other
hand define a simple model for calculating the root and flank temperatures. Generally, the root temperature is considered as
the temperature of the gear body (bulk temperature), and the flank temperature as the current temperature at the contact
point (flash temperature). The flash temperature is very brief and cannot be measured with the used measuring equipment.
In the conducted tests, the gear-body temperature, i.e., the temperature in the tooth root, was measured. According to the
VDI 2736 recommendations, the root temperature is defined with this equation:
 
kϑ ,F uβ Rλ,G
ϑF uβ = ϑ0 + P · μ · HV · + · E D0,64 (6)
b · z · (vt · mn )0,75 AG

A problem that should be highlighted for this material combination is that the temperatures, calculated using Eq. (6),
were lower than the measured values in all the test cases. The temperatures were calculated using the following values:

ϑ0 = 23 ◦ C, b = 6 mm,
P - value is in Table 7, z = 20,
μ = 0.2 for no-lubrication conditions, vt - value is in Table 7,
μ = 0.09 for grease lubrication, mn = 1 mm,
HV = 0.206, Rλ,G = 0,
0,75
K· ( m
s ) · mm1,75
kϑ ,F uβ = 900 W ED = 1.

The discrepancy between the measured and the VDI 2736 model-based temperatures is between 2 and 10%. The dif-
ferences between individual tests for an identical load are within 5%. The discrepancy between the measurements and the
calculations for the tested loads and lubrication conditions is shown in Fig. 18.

Fig. 18. Discrepancy between measured and calculated root temperatures: non-lubricated gear pairs left, lubricated gear pairs right.
D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846 843

Table 8
Experimentally defined values of the coefficient of friction
for the steel/PEEK gear pair in dry and lubricated condi-
tions.

lubrication type average μ standard deviation

dry 0.214 0.019


grease lubrication 0.099 0.008

One of the reasons for the differences between the measurement and the calculations undoubtedly involves the value
of the coefficient kϑ, Fuβ , suggested by the guidelines. It is the same for all polymer materials, meshing with a steel gear in
dry and greased conditions. Deviations from the calculated values also occur because increasing tooth-flank wear results in
increased temperature. After some operating time, the conditions in the contact are no longer the same as at the beginning
of the meshing, and after a sustained period of steady temperature it begins to increase, Fig. 7. The calculated temperature
was compared with the measured temperature just moments before failure, before the temperature soared. If the calculated
temperature was compared with the temperature measured in the earlier test phases when the flanks were not yet worn
out, the discrepancy between the calculation and the measurement would be smaller. The discrepancy between the mea-
sured temperatures in the lubricated case for identical loads was very small, irrespective of the roughness of the steel gear’s
surface. These measurements show that the roughness of the steel gear surface in the lubricated case had no effect on the
operating temperature.
Temperature is closely related to the coefficient of friction between the meshing flanks. The values of the coefficient of
friction are very generally suggested in the VDI 2736 guidelines. For this reason and based on the temperature measure-
ments, in the next step Eq. (6) was used to express the coefficient of friction and determine the coefficient of friction for
the steel/PEEK material combination. The values of the coefficients of friction were determined for dry and lubricated mesh-
ing conditions, Table 8. It would be more accurate to say that the values in Table 8 represent an indicator of the friction
coefficient that is used for the temperature calculation of the gears [19]. The coefficient of friction, in its physical definition,
can vary slightly from these values. The result of calculating the temperature using Eq. (6) depends on the coefficient of
friction, and using the values in Table 8 is suggested for a more accurate temperature calculation of the steel/PEEK gear
pairs. Using the indicator values is more accurate, since it takes into account some hidden influences that determine the
gear temperature, for instance the rolling and sliding ratio in the gear contact, where the rolling friction is much smaller
than the sliding friction. The values presented in Table 8 are mean, since the friction coefficient varies along the path of
contact due to variable normal load, curvature and sliding velocity [70]. The results show that in the case of lubrication, the
two tested surface roughnesses have no significant effect on the coefficient of friction. Within a group of lubricated gear
pairs, the differences between the coefficients of friction, determined for each particular test, are small.

5. Conclusions

The research achieved all the objectives specified at the beginning. The parameters required for a reliable design of PEEK
gears were experimentally defined. The effect of a steel gear surface and the effect of a lubricant on the lifespan of PEEK
gears were tested. Measures to extend the lifespan of polymer gears were proposed. The main conclusions are as follows:
1. Grease lubrication extends the lifespan of a steel/PEEK gear pair. The positive effects of the lubricant are reflected in
slower wear and a reduced coefficient of friction, which results in less heat generation. In the case of ISOFLEX TOPAS NB
52 grease lubrication, the lifespan of the PEEK gears was found to be extended by an average factor of 1.23. Compared
to non-lubricated gear pairs, the measured temperatures of the lubricated gear pairs were more than 20 °C lower.
2. In a steel/PEEK gear-pair meshing, an additional treatment of the steel gear with trovalisation had a significant effect on
the lifespan of the polymer gear. The lifespan of PEEK gears, meshing with trovalised steel gears, increased by an average
factor of 2.54.
3. For two steel-gear surface roughnesses and for lubrication and non-lubrication conditions, the wear coefficients of a
steel/PEEK material pair were determined. The suggested value of the wear coefficient can be used for wear control
according to the VDI 2736 model.
4. Temperature measurements and the VDI 2736 model for calculating the root temperature served as a basis for calculat-
ing the coefficients of friction for the steel/PEEK gear pair in dry and lubricated conditions. The results showed that in
the case of grease lubrication, treating the steel gear’s surface has no significant effect on the coefficient of friction. The
standard deviation of the coefficients of friction, determined within a group of lubricated gear pairs was very small. Dis-
crepancies between the measured and the VDI 2736 model-based temperatures were between 2 and 10%. Discrepancies
between individual tests under identical loads are within 5%. The scatter for identical loads indicates that the calculated
and the measured temperatures are in good correlation.
5. Using numerical simulations, it was found that the size of the tooth-tip fillet radius has a major effect on the flank-
pressure peaks, occurring at the beginning and the end of the meshing. A larger fillet radius results in a lower flank
pressure, which slows down the wear of the polymer gear. By adjusting the flank profile, it is possible to completely
remove the pressure peaks that occur at the beginning and the end of the gear meshing.
844 D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846

6. Based on the presented results and the results of previous work where various materials were tested it can be concluded
that PEEK outperforms the most widely used materials such as POM and PA, also when this basic materials are added
fibres and internal lubricants. Therefore we can state that PEEK is an appropriate material to use for gears which operate
in high demanding conditions (high temperature, high load, or both) and where the material price can be economically
justified.

Acknowledgements

The research was financed partly by the MAPgears project (Slovenian Ministry of Education, Science and Sport, contract
no. C3330-18-952014) and partly by the Slovenian Research Agency (contract no. 630-33/2019-1). The authors would like to
thank the companies Podkrižnik, d.o.o. and Domel, d.d. for manufacturing the test gears and their support.

Supplementary material

Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.mechmachtheory.
2019.07.001.

References

[1] M. Berer, D. Tscharnuter, G. Pinter, Dynamic mechanical response of polyetheretherketone (PEEK) exposed to cyclic loads in the high stress tensile
regime, Int. J. Fatigue 80 (2015) 397–405, doi:10.1016/j.ijfatigue.2015.06.026.
[2] N. Abbasnezhad, A. Khavandi, J. Fitoussi, H. Arabi, M. Shirinbayan, A. Tcharkhtchi, Influence of loading conditions on the overall mechanical behavior
of polyether-ether-ketone (PEEK), Int. J. Fatigue 109 (2018) 83–92, doi:10.1016/j.ijfatigue.2017.12.010.
[3] A. Avanzini, C. Petrogalli, D. Battini, G. Donzella, Influence of micro-notches on the fatigue strength and crack propagation of unfilled and short carbon
fiber reinforced PEEK, Mater. Des. 139 (2018) 447–456, doi:10.1016/j.matdes.2017.11.039.
[4] A.C. Greco, R. Erck, O. Ajayi, G. Fenske, Effect of reinforcement morphology on high-speed sliding friction and wear of PEEK polymers, 18th Int. Conf.
Wear Mater. 271 (2011) 2222–2229, doi:10.1016/j.wear.2011.01.065.
[5] M. Regis, A. Lanzutti, P. Bracco, L. Fedrizzi, Wear behavior of medical grade PEEK and CFR PEEK under dry and bovine serum conditions, Wear 408–409
(2018) 86–95, doi:10.1016/j.wear.2018.05.005.
[6] R. Schroeder, F.W. Torres, C. Binder, A.N. Klein, J.D.B. de Mello, Failure mode in sliding wear of PEEK based composites, Wear Mater 301 (2013) 717–726,
doi:10.1016/j.wear.2012.11.055.
[7] A.S. Mohammed, M.I. Fareed, Improving the friction and wear of poly-ether-etherketone (PEEK) by using thin nano-composite coatings, Wear 364–365
(2016) 154–162, doi:10.1016/j.wear.2016.07.012.
[8] X.-Q. Pei, R. Bennewitz, M. Busse, A.K. Schlarb, Effects of single asperity geometry on friction and wear of PEEK, Wear 304 (2013) 109–117, doi:10.
1016/j.wear.2013.04.032.
[9] V. Rodriguez, J. Sukumaran, A.K. Schlarb, P. De Baets, Reciprocating sliding wear behaviour of PEEK-based hybrid composites, Wear 362–363 (2016)
161–169, doi:10.1016/j.wear.2016.05.024.
[10] V. Rodriguez, J. Sukumaran, A.K. Schlarb, P. De Baets, Influence of solid lubricants on tribological properties of polyetheretherketone (PEEK), Tribol. Int.
103 (2016) 45–57, doi:10.1016/j.triboint.2016.06.037.
[11] R. Shrestha, J. Simsiriwong, N. Shamsaei, Mean strain effects on cyclic deformation and fatigue behavior of polyether ether ketone (PEEK), Polym. Test.
55 (2016) 69–77, doi:10.1016/j.polymertesting.2016.08.002.
[12] R. Shrestha, J. Simsiriwong, N. Shamsaei, R.D. Moser, Cyclic deformation and fatigue behavior of polyether ether ketone (PEEK), Int. J. Fatigue 82 (2016)
411–427, doi:10.1016/j.ijfatigue.2015.08.022.
[13] M. Kurokawa, Y. Uchiyama, S. Nagai, Performance of plastic gear made of carbon fiber reinforced poly-ether-ether-ketone, Tribol. Int. 32 (1999) 491–
497, doi:10.1016/S0301-679X(99)0 0 078-X.
[14] M. Kurokawa, Y. Uchiyama, S. Nagai, Performance of plastic gear made of carbon fiber reinforced poly-ether-ether-ketone: part 2, Tribol. Int. 33 (20 0 0)
715–721, doi:10.1016/S0301-679X(0 0)0 0111-0.
[15] M. Berer, Z. Major, G. Pinter, Elevated pitting wear of injection molded polyetheretherketone (PEEK) rolls, Wear 297 (2013) 1052–1063, doi:10.1016/j.
wear.2012.11.062.
[16] H. Koike, K. Kida, K. Mizobe, X. Shi, S. Oyama, Y. Kashima, Wear of hybrid radial bearings (PEEK ring-PTFE retainer and alumina balls) under dry rolling
contact, Tribol. Int. 90 (2015) 77–83, doi:10.1016/j.triboint.2015.04.007.
[17] D. Kumar, T. Rajmohan, S. Venkatachalapathi, Wear behavior of PEEK matrix Composites: a review, Int. Conf. Adv. Funct. Mater. 5 (2018) 14583–14589
2017 ICAFM’17 03052017 – 05052017, doi:10.1016/j.matpr.2018.03.049.
[18] Verein Deutscher Ingenieure, VDI 2736: Part 2, Thermoplastic gear wheels, Cylindrical gears, Calculation of the load-carrying capacity, VDI-RICHTLIN-
IEN (2014).
[19] J. Tavčar, G. Grkman, J. Duhovnik, Accelerated lifetime testing of reinforced polymer gears, J. Adv. Mech. Des. Syst. Manuf. 12 (2018) JAMDSM0 0 06
–JAMDSM0 0 06, doi:10.1299/jamdsm.2018jamdsm0 0 06.
[20] K. Mao, W. Li, C.J. Hooke, D. Walton, Friction and wear behaviour of acetal and nylon gears, Wear 267 (2009) 639–645, doi:10.1016/j.wear.2008.10.005.
[21] K. Mao, P. Langlois, Z. Hu, K. Alharbi, X. Xu, M. Milson, W. Li, C.J. Hooke, D. Chetwynd, The wear and thermal mechanical contact behaviour of machine
cut polymer gears, Wear 332–333 (2015) 822–826, doi:10.1016/j.wear.2015.01.084.
[22] E. Letzelter, M. Guingand, J.-P. de Vaujany, P. Schlosser, A new experimental approach for measuring thermal behaviour in the case of nylon 6/6
cylindrical gears, Polym. Test. 29 (2010) 1041–1051, doi:10.1016/j.polymertesting.2010.09.002.
[23] M. Kalin, A. Kupec, The dominant effect of temperature on the fatigue behaviour of polymer gears, 21st Int. Conf. Wear Mater. 376–377 (2017) 1339–
1346, doi:10.1016/j.wear.2017.02.003.
[24] J. Moder, F. Grün, M. Stoschka, I. Gódor, A novel two-disc machine for high precision friction assessment, Adv. Tribol. 2017 (2017) 1–16, doi:10.1155/
2017/8901907.
[25] N.A. Wright, S.N. Kukureka, Wear testing and measurement techniques for polymer composite gears, Wear 251 (2001) 1567–1578, doi:10.1016/
S0 043-1648(01)0 0793-1.
[26] A. Pogačnik, J. Tavčar, An accelerated multilevel test and design procedure for polymer gears, Mater. Des. 1980-2015 65 (2015) 961–973, doi:10.1016/j.
matdes.2014.10.016.
D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846 845

[27] J. Duhovnik, D. Zorko, L. Sedej, The effect of the teeth profile shape on polymer gear pair properties, Teh. Vjesn. - Tech. Gaz. (2016) 23, doi:10.17559/
TV-20151028072528.
[28] N. Anand Mohan, S. Senthilvelan, Preliminary bending fatigue performance evaluation of asymmetric composite gears, Mech. Mach. Theory 78 (2014)
92–104, doi:10.1016/j.mechmachtheory.2014.03.006.
[29] C. Hasl, C. Illenberger, P. Oster, T. Tobie, K. Stahl, Potential of oil-lubricated cylindrical plastic gears, J. Adv. Mech. Des. Syst. Manuf. 12 (2018)
JAMDSM0016 –JAMDSM0016, doi:10.1299/jamdsm.2018jamdsm0016.
[30] C. Hasl, H. Liu, P. Oster, T. Tobie, K. Stahl, Forschungsstelle fuer Zahnraeder und Getriebebau (Gear Research Centre), method for calculating the tooth
root stress of plastic spur gears meshing with steel gears under consideration of deflection-induced load sharing, Mech. Mach. Theory 111 (2017)
152–163, doi:10.1016/j.mechmachtheory.2017.01.015.
[31] A.K. Singh, Siddhartha, P.K. Singh, Polymer spur gears behaviors under different loading conditions: a review, Proc. Inst. Mech. Eng. Part J J. Eng. Tribol.
232 (2017) 210–228, doi:10.1177/1350650117711595.
[32] D. Zorko, S. Kulovec, J. Tavčar, J. Duhovnik, Different teeth profile shapes of polymer gears and comparison of their performance, J. Adv. Mech. Des.
Syst. Manuf. 11 (2017) JAMDSM0083 –JAMDSM0083, doi:10.1299/jamdsm.2017jamdsm0083.
[33] G.W. Stachowiak, A.W. Batchelor, Engineering Tribology, 4. ed, BH, Butterworth-Heinemann/Elsevier, Amsterdam, 2014.
[34] J. Moder, F. Grün, F. Summer, M. Kohlhauser, M. Wohlfahrt, Application of high performance composite polymers with steel counterparts in dry
rolling/sliding contacts, Polym. Test. 66 (2018) 371–382, doi:10.1016/j.polymertesting.2018.01.009.
[35] H. Linke, J. Börner, R. Heß, E. Röhle, I. Römhild, M. Senf (Eds.), Cylindrical gears: calculation - materials - manufacturing, Hanser, München, 2016.
[36] D. Mallipeddi, M. Norell, M. Sosa, L. Nyborg, The effect of manufacturing method and running-in load on the surface integrity of efficiency tested
ground, honed and superfinished gears, Tribol. Int. 131 (2019) 277–287, doi:10.1016/j.triboint.2018.10.051.
[37] S. Ramachandra, T.. Ovaert, The effect of controlled surface topographical features on the unlubricated transfer and wear of PEEK, Wear 206 (1997)
94–99, doi:10.1016/S0043-1648(96)07354-1.
[38] D.M. Elliott, J. Fisher, D.T. Clark, Effect of counterface surface roughness and its evolution on the wear and friction of PEEK and PEEK-bonded carbon
fibre composites on stainless steel, Wear 217 (1998) 288–296, doi:10.1016/S0043-1648(98)00148-3.
[39] T.F. de Andrade, H. Wiebeck, A. Sinatora, Effect of surface finishing on friction and wear of Poly-Ether-Ether-Ketone (PEEK) under oil lubrication,
Polímeros 26 (2016) 336–342, doi:10.1590/0104-1428.2183.
[40] M. Andersson, M. Sosa, U. Olofsson, The effect of running-in on the efficiency of superfinished gears, Tribol. Int. 93 (2016) 71–77, doi:10.1016/j.triboint.
2015.08.010.
[41] T.J. Hoskins, K.D. Dearn, Y.K. Chen, S.N. Kukureka, The wear of PEEK in rolling–sliding contact – Simulation of polymer gear applications, Wear 309
(2014) 35–42, doi:10.1016/j.wear.2013.09.014.
[42] A. Avanzini, G. Donzella, A. Mazzù, C. Petrogalli, Wear and rolling contact fatigue of PEEK and PEEK composites, Tribol. Int. 57 (2013) 22–30, doi:10.
1016/j.triboint.2012.07.007.
[43] M. Zalaznik, M. Kalin, S. Novak, G. Jakša, Effect of the type, size and concentration of solid lubricants on the tribological properties of the polymer
PEEK, Wear 364–365 (2016) 31–39, doi:10.1016/j.wear.2016.06.013.
[44] M. Zalaznik, M. Kalin, S. Novak, Influence of the processing temperature on the tribological and mechanical properties of poly-ether-ether-ketone
(PEEK) polymer, Tribol. Int. 94 (2016) 92–97, doi:10.1016/j.triboint.2015.08.016.
[45] M. Kalin, M. Zalaznik, S. Novak, Wear and friction behaviour of poly-ether-ether-ketone (PEEK) filled with graphene, WS 2 and CNT nanoparticles,
Wear 332–333 (2015) 855–862, doi:10.1016/j.wear.2014.12.036.
[46] L. Zhang, H. Qi, G. Li, D. Wang, T. Wang, Q. Wang, G. Zhang, Significantly enhanced wear resistance of PEEK by simply filling with modified graphitic
carbon nitride, Mater. Des. 129 (2017) 192–200, doi:10.1016/j.matdes.2017.05.041.
[47] J.P. Davim, R. Cardoso, Tribological behaviour of the composite PEEK-CF30 at dry sliding against steel using statistical techniques, Mater. Des. 27 (2006)
338–342, doi:10.1016/j.matdes.20 04.11.0 06.
[48] G. Li, H. Qi, G. Zhang, F. Zhao, T. Wang, Q. Wang, Significant friction and wear reduction by assembling two individual PEEK composites with specific
functionalities, Mater. Des. 116 (2017) 152–159, doi:10.1016/j.matdes.2016.11.100.
[49] A.S. Milani, A. Shanian, C. Lynam, T. Scarinci, An application of the analytic network process in multiple criteria material selection, Mater. Des. 44
(2013) 622–632, doi:10.1016/j.matdes.2012.07.057.
[50] D. Miler, D. Žeželj, A. Lončar, K. Vučković, Multi-objective spur gear pair optimization focused on volume and efficiency, Mech. Mach. Theory 125
(2018) 185–195, doi:10.1016/j.mechmachtheory.2018.03.012.
[51] E. Hakimian, A.B. Sulong, Analysis of warpage and shrinkage properties of injection-molded micro gears polymer composites using numerical simula-
tions assisted by the Taguchi method, Mater. Des. 42 (2012) 62–71, doi:10.1016/j.matdes.2012.04.058.
[52] Victrex PEEK 650 G material datasheet. https://www.victrex.com/en/datasheets (accessed 27 May 2019).
[53] DIN 3967:1978-08, System of Gear Fits; Backlash, Tooth Thickness Allowances, Tooth Thickness Tolerances, Principles, Beuth Verlag GmbH, 1978,
doi:10.31030/1103340.
[54] ISO 1328-1:2013, Cylindrical gears - ISO system of flank tolerance classification - Part 1: Definitions and allowable values of deviations relevant to
flanks of gear teeth, 2013.
[55] T. Jabbour, G. Asmar, Stress calculation for plastic helical gears under a real transverse contact ratio, Mech. Mach. Theory 44 (2009) 2236–2247,
doi:10.1016/j.mechmachtheory.20 09.07.0 03.
[56] R.B. Abernethy, The New Weibull handbook: Reliability & Statistical Analysis for Predicting Life, Safety, Survivability, Risk, Cost and Warranty Claims,
4. ed., Abernethy, North Palm Beach, Fla, 2005 print. May 2005.
[57] T.J. Hoskins, K.D. Dearn, S.N. Kukureka, D. Walton, Acoustic noise from polymer gears – A tribological investigation, Mater. Des. 32 (2011) 3509–3515,
doi:10.1016/j.matdes.2011.02.041.
[58] ISO 6336:2006 Parts 1-6, Calculation of load capacity of spur and helical gears, 2006.
[59] DIN 3990:1987, Parts 1-5, Calculation of load capacity of cylindrical gears, Beuth Verlag GmbH (1987).
[60] K. Vučković, I. Galić, Ž. Božić, S. Glodež, Effect of friction in a single-tooth fatigue test, Int. J. Fatigue 114 (2018) 148–158, doi:10.1016/j.ijfatigue.2018.
05.005.
[61] N. Marjanovic, B. Isailovic, V. Marjanovic, Z. Milojevic, M. Blagojevic, M. Bojic, A practical approach to the optimization of gear trains with spur gears,
Mech. Mach. Theory 53 (2012) 1–16, doi:10.1016/j.mechmachtheory.2012.02.004.
[62] Z. Hu, K. Mao, An investigation of misalignment effects on the performance of acetal gears, Tribol. Int. 116 (2017) 394–402, doi:10.1016/j.triboint.2017.
07.029.
[63] D. Miler, A. Lončar, D. Žeželj, Z. Domitran, Influence of profile shift on the spur gear pair optimization, Mech. Mach. Theory 117 (2017) 189–197,
doi:10.1016/j.mechmachtheory.2017.07.001.
[64] P. Garambois, J. Perret-Liaudet, E. Rigaud, NVH robust optimization of gear macro and microgeometries using an efficient tooth contact model, Mech.
Mach. Theory 117 (2017) 78–95, doi:10.1016/j.mechmachtheory.2017.07.008.
[65] M.B. Sánchez, M. Pleguezuelos, J.I. Pedrero, Influence of profile modifications on meshing stiffness, load sharing, and trasnsmission error of involute
spur gears, Mech. Mach. Theory 139 (2019) 506–525, doi:10.1016/j.mechmachtheory.2019.05.014.
[66] V. Roda-Casanova, F. Sanchez-Marin, A 2D finite element based approach to predict the temperature field in polymer spur gear transmissions, Mech.
Mach. Theory 133 (2019) 195–210, doi:10.1016/j.mechmachtheory.2018.11.019.
[67] C.M.C.G. Fernandes, D.M.P. Rocha, R.C. Martins, L. Magalhães, J.H.O. Seabra, Finite element method model to predict bulk and flash temperatures on
polymer gears, Tribol. Int. 120 (2018) 255–268, doi:10.1016/j.triboint.2017.12.027.
846 D. Zorko, S. Kulovec and J. Duhovnik et al. / Mechanism and Machine Theory 140 (2019) 825–846

[68] B. Černe, J. Duhovnik, J. Tavčar, Semi-analytical flash temperature model for thermoplastic polymer spur gears with consideration of linear thermo-
mechanical material characteristics, J. Comput. Des. Eng. (2019), doi:10.1016/j.jcde.2019.03.001.
[69] S. Takanashi, A. Shoji, On the Temperature Risk in the Teeth of Plastic Gears, in: International Power Transmission & Gearing Conference, San Francisco,
1980.
[70] D. Miler, M. Hoić, Z. Domitran, D. Žeželj, Prediction of friction coefficient in dry-lubricated polyoxymethylene spur gear pairs, Mech. Mach. Theory 138
(2019) 205–222, doi:10.1016/j.mechmachtheory.2019.03.040.

You might also like