Vous êtes sur la page 1sur 16

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/245389791

A kinematic analysis of meshing polymer gear teeth

Article  in  Proceedings of the Institution of Mechanical Engineers Part L Journal of Materials Design and Applications · July 2010
DOI: 10.1243/14644207JMDA315

CITATIONS READS

12 602

3 authors, including:

Morad Karimpour Karl D. Dearn


Imperial College London University of Birmingham
15 PUBLICATIONS   114 CITATIONS    69 PUBLICATIONS   688 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Metal Seated Ball Valves for critical application View project

Active tribological control View project

All content following this page was uploaded by Karl D. Dearn on 26 August 2014.

The user has requested enhancement of the downloaded file.


101

A kinematic analysis of meshing polymer gear teeth


M Karimpour1 , K D Dearn2∗ , and D Walton2
1
Mechanics of Materials Division, Department of Mechanical Engineering, Imperial College London,
South Kensington Campus, London, UK
2
Power Transmission Laboratory, School of Mechanical Engineering, The University of Birmingham, Edgbaston,
Birmingham, UK

The manuscript was received on 16 December 2009 and was accepted after revision for publication on 30 April 2010.
DOI: 10.1243/14644207JMDA315

Abstract: This article describes an investigation into the contact behaviour of polymeric gear
transmissions using numerical finite element (FE) and analytical techniques. A polymer gear
pair was modelled and analysed using the ABAQUS software suite and the analytical results were
calculated using the BS ISO 6336 rating standard. Before describing the results, the principles
of the strategies and methods employed in the building of the FE model have been discussed.
The FE model dynamically simulated a range of operating conditions. The simulations showed
that the kinematic behaviour of polymeric gears is substantially different from those predicted
by the classical metal gear theory. Extensions to the path of contact occur at the beginning and
end of the meshing cycle. These are caused by large tooth deflections experienced by polymer
gear teeth, as a result of much lower values of stiffness compared to metallic gears. The prema-
ture contact (occurring at the beginning of the meshing cycle) is hypothesized to be a factor in
pitch line tooth fractures, whereas the extended contact is thought to be a factor in the extreme
wear as seen in experiments. Furthermore, the increase in the path of contact also affects the
induced bending and contact stresses. Simulated values are compared against those predicted
by the international gear standard BS ISO 6336 and are shown to be substantially different. This
is particularly for the case for bending stresses, where analytically derived values are indepen-
dent of contact stiffness. The extreme tooth bending and the differences between analytical and
numerical stresses observed in all the simulations suggest that any future polymeric gear-rating
standard must account for the effects of load sharing (as a result of tooth deflection) and friction
(particularly in dry-running applications).

Keywords: spur gears, polymers, steels, friction, temperature, kinematics

1 INTRODUCTION of employing such polymers in gearing applications


emanate from their low cost, low weight, resilience,
Polymeric materials were first employed in gearing and ability to run dry. A reduced load capacity and,
applications in the 1950s and have developed into crucially, a poor temperature resistance tend to limit
a large range of applications. The majority of these the use of such gears.
tend to be in motion control (low load, temperature, High temperatures reduce the mechanical prop-
and speeds). However, the development of the super erties of polymers much more than metals. One of
engineering polymers and polymer gearing technol- the consequences of this is a reduction in the trans-
ogy has pushed the application limits further into mission accuracy of polymer gears, giving rise to
moderate power transmission functions. The benefits transmission errors caused by the large tooth deflec-
tions. Loss of contact ratio due to the following two
factors should not be ignored in performance specifi-
cation, namely a lower manufacturing accuracy grade
∗ Corresponding author: Power Transmission Laboratory, School of
and an allowance for thermal and hydroscopic expan-
Mechanical Engineering, The University of Birmingham, Edgbas- sion. Klein Meuleman et al. used a quasi-static FE
ton, Birmingham B15 2TT, UK. model to simulate transmission errors in a polymeric
email: k.d.dearn@bham.ac.uk gear set over a range of operational conditions [1].

JMDA315 Proc. IMechE Vol. 224 Part L: J. Materials: Design and Applications
102 M Karimpour, K D Dearn, and D Walton

With this, empirical measurements, and an analytical by one simulation, making it possible to study the
model, they were able to minimize composite tooth- real rolling and sliding contact. Having developed an
to-tooth transmission errors through novel geometric accurate FE model, shaft misalignment and assem-
modifications. bly deflection effects on gear surface durability and
This was not the first time FEA had been used transmission error were also studied. This revealed
to analyse the kinematic behaviour of non-metallic an effective way of reducing transmission error and
gears. Walton et al. used the FE method to study gear surface wear damage through micro-geometry
load-sharing effects and were among the first to modification (i.e. crowning, tip relief, and lead
suggest that, as a result of thermal softening, the correction).
contact ratio of a polymeric gear set was greater
than that predicted by theory [2]. Using a non-
dimensional analysis based on a variety of oper- 2 THEORY
ational parameters, the actual contact conditions
in terms of a contact ratio were determined by The complexities of polymer gearing offer an ideal
a gear elasticity parameter. This work then led to application for the finite-element method. Many
further studies on the beneficial effects of perfor- researchers have used it to study the fundamental
mance and profile modification, such as backlash kinematic and kinetic behaviours of polymer gears. In
allowance [3] and tip relief [4], in non-metallic recent years, as computational power has increased,
gears. along with the sophistication of commercial finite-
The FE method was also employed by Senthilvelan element software, researchers have employed numer-
and Gnanamoorthy to assess the effect of tooth fil- ical techniques not only to study the behaviour of
let radius on gear performance [5]. They employed a polymeric gear transmissions, but also to modify and
basic single-tooth model, loaded only at the tip, and optimize gear trains for specific applications. However,
disregarded contact ratio effects, calculating an equiv- these polymer gear simulations are all based on quasi-
alent line load. This indicated, as would be expected, static solutions (as detailed above). The following
higher bending stress levels in teeth with smaller fillet section discusses current rating methods for non-
radii. There were similar increases in tooth deflec- metallic gears and discusses those that are deemed
tions, having the combined effect of shortening the most appropriate to the unique behaviour of plastic
gear life. gears. The development of a dynamic non-linear FE
Van Melick used both FEM and analytical methods model to study the kinematic behaviour of a poly-
to investigate the influence of stiffness on dissimilar mer gear transmission is then described. Employing
materials, gear kinematics, and stresses (a steel gear a dynamic solution allowed the whole meshing cycle
is typically 70 times stiffer than an equivalent poly- to be continuously simulated, resulting in a better
oxymethylene polymer gear) [6]. He suggested that understanding of tooth bending effects, and the rami-
polymeric gear kinematics are different to those pro- fications of these effects for other aspects of polymer
duced by the classical gear theory, with the effect of gear performance. In addition to this, it provided an
increasing stresses and, through an extended path of answer as to whether gear-rating standards, developed
contact, influencing the wear resistance of the plastic specifically for metallic gears, could be used to design
gear. An interesting aspect of this study was the discov- and rate polymer gears.
ery of a reciprocating motion at the root of the driven
gear as the teeth disengage at the end of the meshing
cycle. The FEA model used employed a quasi-static 2.1 Design and rating standards
solution.
FEA has been used for some time to model the com- Gear-rating procedures adopt different assumptions
plexities of the steel gear theory. One of the most to give the best approximation of stresses within and
detailed and accurate FEA simulation on metallic around gear teeth; hence, it is inevitable that differ-
gears was arguably conducted by Mao [7]. He utilized ent procedures will predict different values according
FE to investigate the effects of micro-geometry modi- to the assumptions made. In a comparison of avail-
fications on the reduction of transmission errors and able procedures for the rating of non-metallic gears
fatigue damage in a metallic helical gear set. To achieve (BS6168, ESDU68001, and Polypenco), Walton and Shi
the maximum possible geometrical accuracy, instead observed large discrepancies between different meth-
of importing a model into the FE software (ABAQUS), ods and suggested that an experimental investigation
the gear geometry was mathematically generated was required to assess which rating standard was the
using Python Script. Advanced surface-based tie tech- most accurate [8]. Following on from this, Cropper
niques between nodes were utilized to obtain a high- compared the stress levels predicted by a similar range
quality mesh for contact. The novel aspect of this was of rating standards against an FE analysis considering
that instead of a multi-simulation technique (quasi- load sharing and showed BS ISO 6336 (method B) to
static), the whole gear meshing process was achieved be the most accurate [9].

Proc. IMechE Vol. 224 Part L: J. Materials: Design and Applications JMDA315
Kinematic analysis of meshing polymer gear teeth 103

2.2 Bending stress estimation


All bending stress analyses are based on the Lewis
bending equation, which has gone through many
minor modifications since its first introduction. For
instance, ISO 6336:1996 [10] has adopted the exact
location of maximum stress experimentally found by
Hoffer [11]. ISO 6336 (methods B and C) as well as
ANSI/AGMA 2001 take into account the stress concen-
tration at the tooth root fillet, which is ignored by the
other methods. A unique characteristic of method B in
ISO 6336 is that it enables the tooth form factor to be
calculated from first principles (an extensive process,
but it accounts for any tooth geometry); other methods
use tabulated data instead.

2.3 Contact stresses


The basis of all contact stress calculations is the classi-
cal Hertzian contact analysis. Stress levels are usually
calculated at the pitch point where the sliding fric- Fig. 1 Geometric specification of the ‘Birmingham
tion may be assumed to be zero as well as negligible Benchmark’ gear geometry
bending deflection due to contact. The only differ-
ence between the standards is the load-sharing effect.
ANSI/AGMA 2001 assumes no load sharing, whereas
ISO 6336 does consider a load-sharing factor in the to the accuracy of the solution. The size of the ele-
calculation of contact stress. When considering poly- ments was determined based on the simulation of
meric materials, load sharing cannot be disregarded metallic gear teeth from reference [12]. As the polymer
and hence ISO 6336:1996 would seem to be the most material used in the simulations has a lower stiff-
accurate available method. When comparisons are ness than the material used by Wang and Howard, the
made against a standard in this article, they are made field variable gradients are much smaller than those of
against ISO 6336:1996 method B (from this point metallic gears. Hence, the same mesh density with a
forward referred to as the Standard). structured mesh throughout the tooth has been used.
This will be conservative in terms of mesh refinement
but should guarantee accurate results. The rest of the
3 FINITE-ELEMENT MODEL FOR PLASTIC GEAR teeth were modelled with a coarse mesh to increase
TOOTH KINEMATIC ANALYSIS computational efficiency.
As the stress elements do not have rotational degrees
All gear theories as well as the gear-rating standards of freedom, an advanced surface-based tie technique
are based on rigid-body kinematics. This assumes that was utilized to attach the nodes at the gear hub
no significant deformation occurs in the tooth, pre- to a rigid-body shaft, later used to apply rotational
serving the involute profile. This is probably a valid loads and boundary conditions (also shown in Fig. 2).
assumption in the metallic gear theory; however, it is In order to minimize fluctuations between contact-
unlikely to be the case for plastic gears in all but the ing surfaces, an inertia load was assigned to the
lightest of applications. shaft to dampen unwanted vibrations. The elements
An accurate two-dimensional model of the Birm- used (CPE4R) were four-node bilinear plain strain
ingham benchmark gear was developed and imported quadrilaterals with reduced integration [13].
into a state-of-the-art FE package (ABAQUS). A sum- Solutions were obtained in three steps, namely
mary of the gear geometry is given in Fig. 1. In order approach, loading, and rotation, selected to reduce
to reduce the computational time of the simulation, the occurrence of contact inaccuracies and vibration.
only ten meshing teeth were modelled, with two pairs A smooth step profile was defined as custom ampli-
of mating teeth meshed to a higher mesh density tude to apply all the boundary conditions and loads
to enable the extraction of accurate data (shown in as smoothly as possible. Boundary conditions and
Fig. 2). The mesh in this region has relatively small loads were applied at the drive shaft reference points
elements dimensions (i.e. approximately 0.04 mm). (centres), according to the solution step sequence.
In addition to this, the use of a structured mesh of Four separate surfaces were defined on each mesh-
quadrilateral elements following the involute profile ing gear, to isolate each zone of the meshing cycle. The
of the teeth and enhanced hour-glass control adds boundary conditions were selected as follows.

JMDA315 Proc. IMechE Vol. 224 Part L: J. Materials: Design and Applications
104 M Karimpour, K D Dearn, and D Walton

Fig. 2 (a) Loads and boundary conditions and (b) generated mesh representing the gear

1. Initial step: Boundary conditions were applied to Table 1 Simulation schedule


constrain completely the driving and driven gears.
E (GPa) T (N m) μ Notes
2. Approach: A rotational constraint was removed
from the driving gear, then a rotational boundary Test 1 3.1 5 0
condition was activated, closing the ‘gap’ between Test 2 3.1 8 0
Test 3 3.1 11 0
interacting tooth flanks. The magnitude of rotation Test 4 3.1 15 0
was equivalent to the theoretical backlash for the Test 5 2.5 8 0 Analogous to POM (Delrin® 500)
benchmark geometry. at 35 ◦ C
Test 6 2 8 0 Analogous to POM (Delrin® 500)
3. Loading: A rotational constraint was removed from at 50 ◦ C
the driven gear (thus both gears were now free to Test 7 1 8 0 Analogous to POM (Delrin® 500)
rotate). Opposing rotational moments were then at 90 ◦ C
Test 8 3.1 8 0.3
applied on the driving and driven gears, essentially Test 9 3.1 8 0.5
resisting each other.
4. Rotation: Rotational boundary conditions were
applied about the drive shaft centres. The magni- polymer manufacturer Du Pont [15]. For each simula-
tude of the rotations allowed at least one pair of tion, Poisson’s ratio was taken as 0.35 and the material
gear teeth to complete one meshing cycle. model used was isotropic linear elastic.

Simulations were conducted for combinations


shown in Table 1. Load-sharing data were smoothed 4 RESULTS
using the Stavitzky–Golay [14] method implemented
within MATLAB. This was to compensate for the fluc- With the FE model completely defined, simulation
tuations caused by the inevitable numerical errors and analysis were initiated to test a variety of com-
and contact inaccuracies (this was a result of the way mon operational conditions for polymer gears. These
the contact algorithm is defined in ABAQUS). Mate- were specifically selected to examine the explicit kine-
rial properties for the simulations were taken from the matic responses and to assess the suitability of using

Proc. IMechE Vol. 224 Part L: J. Materials: Design and Applications JMDA315
Kinematic analysis of meshing polymer gear teeth 105

Fig. 4 A schematic of the premature contact caused by


large tooth deflections

Fig. 3 Extensions to the path of contact extracted from governed by the mechanical properties of the polymer.
the FE model simulating an applied load of 7 N m This can be verified by the results presented by Wang
(E = 3.1 GPa) and Howard [12]. It was shown that in soft metallic
gears (i.e. aluminium gears), the load share ratio pro-
file differs from the theoretical curve and that with
the Standard for the rating of polymeric gears. Stresses greater load, the extent of premature and extended
were then examined and compared against those contact would be increased.
predicted by the Standard, highlighting the shortcom- During the premature contact stage, the tip of the
ings of this rating method. Finally, the effects that tooth and top land of the wheel make heavy contact
operational conditions (for example, temperature and with the pinion tooth in the proximity of its pitch point.
friction) had on the performance of the gears are given. Contact then continues up the tooth flanks until they
are tangent to one another, followed by ‘normal’ con-
tact until the theoretical LPC. This was likely to be the
4.1 Kinematics result of a deflection ‘lag’ (i.e. it was the result of suc-
cessive teeth already being deflected). After this point,
4.1.1 Path of contact
the deflected teeth attempt to return to their original
The classical gear theory predicts that the path of form. In doing so, the tip of the pinion slides along
contact of a pair of meshing gears lies on a straight the flank of the wheel in the direction of the pitch
line (effectively being the locus of all contact points) point. This was the reciprocating motion that was first
between the first (FPC) and last points of tooth con- suggested by van Melick and is discussed below.
tact (LPC) (governed by the addendum radii). Figure 3 Van Melick [6] suggested that the reciprocation of
shows the locus of contact points extracted from the tooth tip at the mating root accounts for the dis-
simulation 2 (specified in Table 1), plotted against tinctive wear patterns described by Breeds et al. [16],
the theoretical benchmark geometry. It exhibits, as such that they govern the mechanisms of wear in poly-
expected, a line of contact between the addendum meric gearing. This is a strong statement given that
radii as predicted by the theory. However, there are Walton et al. [3] reported that wear was the predomi-
also periods of premature contact (Fig. 4) occurring nant form of failure in polymer gears. Given the locality
before the theoretical FPC and extended contact after of the premature contact region (i.e. at the pitch line
the predicted LPC. The extensions do not coincide with of the pinion), this may also be a contributory factor
the linear section of the line but are almost perpen- in pitch line fractures (PLFs). This is particularly inter-
dicular to it. The extended contact (at the LPC) lies esting considering the work of Cropper [9], who noted
almost exactly on the addendum radius of the wheel, that for the benchmark geometry (i.e. when the pinion
suggesting that this extraordinary contact occurs on and wheel are geometrically identical) PLF occurs only
the involute area of the tooth flank. This was not the on the pinion. This, however, requires further research.
case for the premature contact region, suggesting that
contact occurs outside the involute flank on the tooth 4.1.2 Load sharing
tip.
The main reason for this behaviour lies with the The fraction of the applied load transmitted by individ-
very large tooth deflections observed in polymeric gear ual gear teeth was governed by the theoretical contact
teeth, deformations that are much greater than those ratio, which, in case of the benchmark geometry, was
found in metallic gearing. These deflections signifi- 1.65. Physically, this means that for approximately
cantly alter the (theoretical) tooth geometry and are two-thirds of the meshing cycle, two pairs of teeth

JMDA315 Proc. IMechE Vol. 224 Part L: J. Materials: Design and Applications
106 M Karimpour, K D Dearn, and D Walton

Fig. 5 The load share ratio against the roll angle predicted by simulation 2

carry the load, whereas for the remainder a single pair


carries it. The division of the applied load between dif-
ferent meshing tooth pairs is defined as the load share
ratio. The Standard defines the load share fraction (for
metal gears) as 1/3, 2/3, and 3/3 (i.e. at the first point
of contact one-third of the load is carried by a mesh-
ing pair and this increases to two-thirds at the point of
single tooth contact where finally the entire load is car-
ried). The reverse of this then occurs as the tooth pair
runs out of the mesh.
Figure 5 shows the load share ratio plotted against
the roll angle in simulation 2. It reaches a maximum of
approximately 0.8, suggesting that single tooth con-
tact does not occur in polymeric gears under such
conditions, pushing the real contact ratio above 2.
This clearly invalidates the load sharing predicted by
the Standard. The parabolic form of the load share
ratio also suggests that deflections are shared evenly
between the meshing pair, which is in agreement with
the work of Klein Meuleman et al. [1]. With the max-
imum load occurring at the pitch line of the teeth, it
is in this area where maximum contact and bending
stresses are expected. Another interesting aspect of
Fig. 5 is the increase in the roll angle (the rotational
angle of gear body from FPC to LPC). Theoretically,
the roll angle of the benchmark geometry should be
19.84◦ ; however, this value increases in the simulations
by 30 per cent to 25◦ . The reason for this is the cumula-
tive effect of the extensions to the path of contact and Fig. 6 Position of nodes used to locate the position of
ultimately the large tooth deflections. maximum: (a) bending and (b) contact stresses

4.2 Stress analysis


a rating for it as a gear material (i.e. Von Mises yield
In the following sections, Von Mises stress was chosen criteria). The nodes from which the predicted stresses
as the scalar representative of the stress state at a given were extracted from the tooth flank and root are shown
point. At the tooth root, this would translate into the in Fig. 6.
magnitude of tensile stresses caused by bending. In
the case of the contact patch, this would represent the 4.2.1 Maximum bending stress
stresses caused by contact between the mating tooth
flanks. As in the Standard, this could then be com- The extracted bending stress predicted by simulation 2
pared with the limits of the material in order to provide for each node is plotted against the roll angle in Fig. 7.

Proc. IMechE Vol. 224 Part L: J. Materials: Design and Applications JMDA315
Kinematic analysis of meshing polymer gear teeth 107

Fig. 7 Bending stresses at various nodes against the roll angle predicted by simulation 2

It takes a similar form to the load-sharing graph above; assumes a bending stress maximum at tip loading
it also shows the location of the nodes from which data (being the point at which the lever arm is greatest);
were taken. The maximum stress occurs at the ‘centre’ however, it does not consider that at this point, the load
of the fillet radius (node 2539) and when the point of is being carried by (at least) two teeth pairs reducing
contact is at the pitch point. This is not, however, the the tangential force. The effect of load sharing reduces
point at which the Standard predicts that maximum the magnitude of the applied load, with the peak at the
stress will occur. The Standard predicts that the great- pitch point, where maximum stress occurs.
est stress is induced at the tooth tip (i.e. the first or last
point of contact).
4.2.2 Maximum contact stress
The difference between the assumption made in
the Standard and the location of maximum stress pre- A similar approach was adopted to establish the posi-
dicted in the simulation has a significant effect on the tion of maximum contact stress across the tooth flank.
predicted bending stress values. The simulation pre- Figure 8 shows the corresponding stress profile plot-
dicts a maximum bending stress of 25.9 MPa, whereas ted against the roll angle along with the nodes along
the Standard calculates it to be 38 MPa. This is a signifi- the tooth flank. Once again, the maximum stress value
cant difference of around 46 per cent. Tooth deflection occurred at the pitch point. A maximum contact stress
is the most likely explanation for this discrepancy. value of 38.69 MPa was predicted. This is in close
The Lewis equation, on which the Standard is based, agreement with the Standard that predicts a value of

Fig. 8 Contact stresses at various nodes against the roll angle predicted by simulation 2

JMDA315 Proc. IMechE Vol. 224 Part L: J. Materials: Design and Applications
108 M Karimpour, K D Dearn, and D Walton

Fig. 9 Induced stress contours during meshing simulations 2 (left) and 4 (right)

contact stress of 37.57 MPa (a difference of 2.9 per (representing initial contact when the tip/top land of
cent). the wheel tooth collides with the pitch point of the
Caution should, however, be exercised when using pinion tooth). Extended contact is apparent in the
the Standard to calculate contact stresses. The increase in the stresses seen in node 2093.
single-pair tooth contact factor, used in the Standard,
converts the contact stress calculated at the pitch point 4.2.3 Effect of applied loads
to that at the inner point of contact [17]. This may go
some way to explaining the discrepancy but certainly The analysis conducted to assess the effect of load on
requires further examination. The stress contours with the contact and bending stresses revealed some inter-
the gear teeth during loading, during simulations 2 and esting results. Four separate simulations were carried
4, are shown in Fig. 9. out to simulate a range of applied loads. A compar-
It is also interesting to note the manifestation of the ison of the stresses generated from the simulations
extended contact points in Fig. 8. Evidence of prema- and those calculated using the Standard is given in
ture contact appears in the double peak seen in the Table 2. The values of contact stress found using the
plot of node 2139 and the single peak in node 2143 simulations correlate well with the Standard’s values.

Proc. IMechE Vol. 224 Part L: J. Materials: Design and Applications JMDA315
Kinematic analysis of meshing polymer gear teeth 109

Table 2 A comparison of FE and BS ISO 6336-derived


stresses for a range of loads

Contact stress (MPa) Bending stress (MPa)


Torque FE BS ISO FE BS ISO
(N m) simulation 6336 simulation 6336

Test 1 5 31.57 29.71 17.13 23.80


Test 2 7 38.70 37.58 25.94 38.08
Test 3 10 44.20 44.06 34.36 52.36
Test 4 15 48.75 49.71 42.31 66.64

This is not the case for the bending stresses however,


where discrepancies between simulated and the Stan-
dard values increase with load. This makes the Stan-
dard (part 3) even less suitable for rating polymer gears
in demanding applications. Fig. 11 Effect of applied loading on the path of contact
These stresses should be carefully interpreted, as
they are those that occur at the pitch point. It has,
however, already been shown that because of extended initial contact occurs closer to the pitch point) and that
contact (specifically the interference of tooth tip and extended contact lasts longer. Thus, the overall length
mating flank), high contact stress peaks occur during of the contact path is increased.
the early and later stages of the meshing cycle. The
stresses shown in Table 2 do not account for these.
Figure 10 shows the kinematic effect of increasing 4.3 Effects of operating conditions
the loads on meshing gear teeth. It indicates that the 4.3.1 Temperature
maximum load carried by a single gear tooth pair
decreases as the applied load increases. This implies The effect of temperature in this case is taken to rep-
that the higher loads induce greater tooth deflections, resent the reduction in the material stiffness because
increasing the real contact ratio and load sharing. of elevated operating temperatures. This is a simpli-
There are other kinematic effects induced by these fication of the physical manifestation of high gear
deflections, namely larger extensions to the path of the temperatures; however, in terms of the kinematic
contact. Figure 11 compares the paths of contact for behaviour of the gear teeth, the reduction in modulus
each of the load conditions specified in Table 1. This has a strong influence. Table 3 summarizes the reduc-
suggests that greater tooth deflections cause prema- tion in stiffness with temperature rise. These values are
ture contact to occur earlier (it therefore follows that assigned to the materials specified in simulations 2, 5,

Fig. 10 Effect of applied loading (torque) on load sharing

JMDA315 Proc. IMechE Vol. 224 Part L: J. Materials: Design and Applications
110 M Karimpour, K D Dearn, and D Walton

Table 3 Material properties as a function of gear body A decrease in stiffness increases the level of load
temperature sharing experienced by individual gear teeth (Fig. 12)
Equivalent
and has a similar effect on the path of contact. The
Modulus of gear body influence of temperature on kinematic behaviour is
elasticity Torque temperature similar to the effects of increasing load, as shown in
(GPa) (N m) (◦ C)
Fig. 10.
Test 2 3.1 7 23 A comparison of simulated and Standard derived
Test 5 2.5 7 35 bending and contact stresses is shown in Fig. 13.
Test 6 2 7 50 Once again, contact stresses correlate well with one
Test 7 1 7 90
another and decrease with a reduction in stiffness
Material data taken from reference [13]. (an effect that was expected as the Standard derived
contact stresses account for contact stiffness). This
6, and 7 (all transmit a load of 7 N m), which are based has implications in terms of gear operating perfor-
on the homo-polymer polyoxymethylene as specified mance, particularly at elevated temperatures. This
by Du Pont [15]. could explain why Kono [18] observed an increase in

Fig. 12 Effect of increasing temperature (i.e. the reduction in modulus) on load sharing

Fig. 13 Comparison of contact and bending stresses versus Young’s modulus, at the pitch point

Proc. IMechE Vol. 224 Part L: J. Materials: Design and Applications JMDA315
Kinematic analysis of meshing polymer gear teeth 111

Fig. 14 Effect of friction on load sharing

Fig. 15 Comparison of contact and bending stresses versus coefficient of friction

gear durability during his experimental programme at contact and bending stresses of polymer gears. Yet,
150 ◦ C. Interestingly, bending stresses do not correlate Walton et al. [19] have shown that coefficients of fric-
well. FE-determined bending stresses increase with tion in dry-running polymeric gears can be as high as
stiffness, whereas the Standard derived stresses are 0.8. Frictional forces induced by coefficients of fric-
overestimated and remain constant. This is because tion of this magnitude must affect the position and
the Standard bending stress is, in theory, dependent magnitude of contact stresses and to a lesser extent
only on geometry and load. For BS ISO 3663 to become the stresses in bending. Simulations 2, 8, and 9 were
a better approximation of polymeric gear bending executed, transmitting a load of 7 N m with increas-
stresses, it should incorporate a material factor that ing coefficients of friction. These were conducted to
would account for load sharing due to tooth bending. investigate the effect of increasing friction on the stress
and kinematic behaviour of the gears.
4.3.2 Friction The gear theory dictates that at the pitch point, the
gears experience only a rolling action (with no slid-
The Standard does not consider the effects of friction, ing velocity) and so assumes a negligible frictional
because it is written for metallic gears that are always influence. Figure 14 suggests that for (dry-running)
lubricated and where frictional forces are low. No con- polymeric gears this is not the case. It shows that
sideration is given to the effect friction will have on frictional forces decrease the maximum load share

JMDA315 Proc. IMechE Vol. 224 Part L: J. Materials: Design and Applications
112 M Karimpour, K D Dearn, and D Walton

ratio and hence increases transmitted loads and sub- 5. Extended contact and van Melicks’ suggested
sequent contact stresses. Friction does not seem to reciprocating motion at the tooth root for
affect the kinematic behaviour of the gears and has polymer–steel contacts were confirmed in the case
a negligible effect on the length of the path of contact. of all the polymer gear meshes simulated here.
Similar trends, shown in Fig. 15, are observed when 6. The extensions to the path of contact increased the
FE and the Standard derived stresses are compared roll angles predicted by the theory.
against increasing coefficients of friction. As expected, 7. For the conditions simulated, load-sharing ratios
Standard derived stresses show no variation with fric- were always below 1. This implied that the real
tion. The FE simulations increasingly overestimate contact ratios were always above 2, despite the
contact stresses compared with those calculated from theoretical prediction of 1.65.
the Standard. FE contact stresses show a strong depen- 8. As the Standard does not consider tooth deflec-
dence on friction. Bending stresses are also shown to tions, reducing the Young’s modulus of the
increase with friction but remain below stress levels polymer (as would be the case at elevated oper-
predicted by the Standard. ational temperatures) has no effect on Standard-
Therefore, for the Standard to become a more accu- predicted bending stresses. It does, however, with
rate means of specifying polymeric gears, a frictional the FE-derived bending stresses. Increasing the
factor should be incorporated into the rating equa- stiffness of the modelled gears is likely to cause
tions to account for tangential forces, particularly in the simulated bending stresses to converge with
dry-running applications. The dominance of contact those of the Standard.
stresses (being greater than the bending stresses) may 9. Frictional effects cannot be disregarded in dry-
also contribute to wear being the predominant failure running polymer gears where they were shown
mechanism in dry-running polymeric gears. to have a significant effect on the induced tooth
stresses.
10. Loss of contact ratio due to allowance for expan-
sion and lower tooth accuracy need to be consid-
5 CONCLUSIONS ered.

This article has simulated the kinematic and kinetic The above conclusions point to the need for a spe-
behaviour of a pair of dry-running, similar-material, cific polymer gear Standard that accounts for the
non-metallic gears running under a variety of idiosyncrasies that are not based on metallic gear-
operating conditions. The stresses extracted from the rating methods. The authors have evidence to show
finite-element models have been compared to those that the lack of applicable design data and a rating
calculated from the BS ISO 6336 rating Standard Standard, in the public domain, is preventing this
(method B). The main conclusions are as follows. novel form of gearing from being fully exploited.

© Authors 2010
1. The assumptions made by the classical gear the-
ory and inherited by most common gear-rating
standards, specifically those of negligible tooth REFERENCES
deflections and frictional effects, are not valid for
dry-running non-metallic gears that have high 1 Klein Meuleman, P.,Walton, D., Dearn, K. D.,Weale, D. J.,
and Driessen, I. Minimization of transmission errors in
friction coefficients.
highly loaded plastic gear trains. Proc. IMechE, Part C:
2. Employing a non-linear FE technique to develop a
J. Mechanical Engineering Science, 2007, 221(C9), 1117–
dynamic simulation of the meshing cycle enabled 1129. DOI: 10.1243/09544062JMES439.
the entire kinematic history of the cycle to be 2 Walton, D., Tessema, A. A., Hooke, C. J., and Shippen,
recorded (including the load-sharing effects and J. M. Load sharing in metallic and non-metallic gears.
increases to the path of contact) and also allowed Proc. IMechE, Part C: J. Mechanical Engineering Science,
the frictional effects to be investigated during 1994, 208(C2), 81–87. DOI: 10.1243/PIME_PROC_1994_
contact. These were shown to have a significant 208_104_02.
influence on FE-derived stresses. 3 Walton, D., Tessema, A. A., Hooke, C. J., and Shippen, J.
3. Contact is shown to occur outside the theoreti- M. A note on tip relief and backlash allowances in non-
cal line of contact, increasing the roll angle of the metallic gears. Proc. IMechE, Part C: J. Mechanical Engi-
neering Science, 1995, 209(C6), 383–388. DOI: 10.1243/
gears.
PIME_PROC_1995_209_169_02.
4. Premature contact extends the beginning of the
4 White, J.,Walton, D., and Weale, D. J. The beneficial effect
contact line; the physical result of this is an impact of tip relief on plastic spur gears. In Proceedings of the
between the tip of the driven gear and the region Conference at ANTEC’98, Society of Plastics Engineers,
of the pitch point of the driver gear. This induces Atlanta, USA, April 1998, vol. III, pp. 3013–3017.
abnormally high levels of contact stress and could 5 Senthilvelan, S. and Gnanamoorthy, R. Effect of gear
be a contributory factor in PLFs. tooth fillet radius on the performance of injection

Proc. IMechE Vol. 224 Part L: J. Materials: Design and Applications JMDA315
Kinematic analysis of meshing polymer gear teeth 113

moulded nylon 6/6 gears. Mater. Des., 2006, 27(8), 632– Subscripts
639.
6 van Melick, H. G. Influence of Young’s modulus on kine- 1 pinion
matics and stresses in plastic spur gears. In Proceedings 2 wheel
of the Conference at VDI Breichte, 1904, 2005, vol. II,
pp. 1219–1225.
7 Mao, K. An approach for power train gear transmis- APPENDIX 2. BS ISO 6336 (METHOD B)
sion error prediction using the non-linear finite ele-
ment method. Proc. IMechE, Part D: J. Automotive A2.1 Part 2: calculation of surface durability
Engineering, 2006, 220(D10), 1455–1463. DOI: 10.1243/
09544070JAUTO251. Surface durability is assessed according to nominal
8 Walton, D. and Shi, Y. W. A comparison of ratings for contact stress levels calculated at the pitch point or
plastic gears. Proc. IMechE, Part C: J. Mechanical Engi- those at the inner point of single tooth contact. The
neering Science, 1989, 203(C1), 31–38. DOI: 10.1243/ nominal contact stress, σH , and the permissible con-
PIME_PROC_1989_203_083_02. tact stress, σHP , can be calculated using equations (1)
9 Cropper, A. B. Failure mode analysis of polymer gears. and (2), respectively.
PhD Thesis, University of Birmingham, UK, 2003. 
10 BS ISO 6336. Calculation of load capacity of spur σH = ZB σHO KA KV KH α KH β (1)
and helical gears, Part 3: calculation of tooth bending 
strength, 2nd edition, 1996 (British Standards Institution, Ft u + 1
σHO = ZH ZE Zβ Zε (2)
London). D1 b u
11 Hoffer, H. Verzahnungskorrekturn an Zahnraden (tooth
corrections to gear wheels). Automob. Z., 1947, 49(2), A2.1.1 ZH : zone factor for contact stress
19–20.
12 Wang, J. and Howard, I. The torsional stiffness of The zone factor accounts for the influence of tooth
involute spur gears. Proc. IMechE, Part C: J. Mechan- flank curvature at the pitch point and converts tan-
ical Engineering Science, 2004, 218(C1), 131–142. DOI: gential force to a normal force at the pitch point
10.1243/095440604322787009. (equation (3))
13 Dassault Systèmes. Simulia, 2007 (Warrington, UK). 
14 Bromba, M. U. A. and Ziegler, H. Application hints for 2 cos βb cos αwt
Savitzky–Golay digital smoothing filters. Anal. Chem., ZH = (3)
1981, 53, 1583–1586. cos2 αt sin αwt
15 Du Pont PLC. Engineering polymers for gearing applica-
tions, 1998 (Du Pont PLC, Bristol). A2.1.2 ZE : elasticity factor for contact stress
16 Breeds, A. R., Kukureka, S. N., Mao, K., Walton, D., and
Hooke, C. J. Wear behaviour of acetal gear pairs. Wear, The elasticity factor accounts for the influence of
1993, 66, 85–91. material properties (equation (4)).
17 BS ISO 6336. Calculation of load capacity of spur and 
helical gears, Part 2: calculation of surface durability (pit- 1
ZE = (4)
ting), 2nd edition, 1996 (British Standards Institution, π{[(1 − v12 )/E1 ] + [(1 − v22 )/E2 ]}
London).
18 Kono, S. Increase in power density of plastic gears for A2.1.3 Zε : contact ratio factor
automotive applications. PhD Thesis, University of Birm-
ingham, UK, 2003. The contact ratio factor, Zε , considers the influence of
19 Walton, D., Cropper, A. B.,Weale, D. J., and Klein Meule- the transverse contact and overlap ratios on the tooth
man, P. The efficiency and friction of plastic cylindrical surface load capacity (equation (5)).
gears. Part 1: influence of materials. Proc. IMechE, Part For spur gears
J: J. Engineering Tribology, 2002, 216(J2), 75–78. DOI: 
10.1243/1350650021543915. 4 − εα
20 BS 6168: 1987. Specification for non-metallic spur-gears, Zε = (5)
3
1987 (British Standards Institution, London).
The transverse contact ratio, εα , is given in
APPENDIX 1 equation (6) and the overlap ratio in equation (7).
Transverse base pitch
Notation
Pbt = mt π cos αt
E Young’s modulus (GPa)
ra addendum radius Length of path of contact
T torque (N m)   
1
gα = da1 − db1 ± da1 − db1 − a sin αwt
2 2 2 2

μ coefficient of friction 2
ϕ theoretical pressure angle (degree) (positive sign for external gears)

JMDA315 Proc. IMechE Vol. 224 Part L: J. Materials: Design and Applications
114 M Karimpour, K D Dearn, and D Walton

Transverse contact ratio Calculation of hFe


gα d n = mn z n dbn = dn cos αn
εα = (6)
Pbt εα
dan = dn + da − d εαn =
cos2 βb
Transverse base pitch  ⎡
   
dan 2 dbn 2

b sin β −
εβ = (7) 2 2
πmn z
den = 2 ⎤2
|z|  
πd cos β cos αn dbn 2
A2.1.4 ZB and ZD : single-pair tooth contact factors
− (εαn − 1)⎦ +
|z| 2
The single pair tooth contact factors transform the  
contact stress at the pitch point to the inner point dbn
αen = cos−1
of single pair tooth contact ((if ZB > 1 or ZD > 1), den
equation (8)).
0.5π + 2 tan αn x
γe = + inv αn − inv αen
tan αwt zn
M1 =  
αFen = αen − γe

2
(da1 /db12
) − 1 − (2π/z1 )
π 
 hfe cos αn

= 0.5zn − cos −θ
× 2
(da2 2
/db2 ) − 1 − (εα − 1)(2π/z2 ) mn cos αFen 3
 
ρfp G
tan αwt + 0.5 −
M2 =  mn cos θ


(da2 /db2 ) − 1 − (2π/z2 )
2 2
Calculation of tooth root normal chord



SFn π  √  G ρfp

× 2
(da1 2
/db1 ) − 1 − (εα − 1)(2π/z1 ) = zn sin −θ + 3 − (12)
mn 3 cos θ mn
ZB = 1 if M1  1 ZD = 1 if M2  1 An accurate value of θ is required to evaluate hFe ,
SFn , and αFen ; it can be calculated through iteration
ZB = M1 if M1 > 1 Z D = M2 if M2 > 1
using equation (13) and an iterative seed value of π/6
(8) (suggested by BS 6168: 1987 [20]).
Calculation of θ

A2.2 Part 3: calculation of tooth bending strength π ρfp


E= mn − hfp tan αn − (1 − sin αn )
4 cos αn
Nominal tooth bending strength is calculated at the
tooth root using a modified form of the Lewis bending where ρfp is the basic rack root fillet radius, Spr the
equation (given in equation (9)). In a similar manner to undercut factor (set to zero for polymer gears), and hfp
that outlined above, multiplying the nominal bending the dedendum of the basic rack
stress by a series of correctional factors gives a permis-
sible stress level (equation (10)). Again, using ISO 6336 ρfp hfp
G= − +x
for polymeric gears, most of the correctional factors mn mn
were set to unity
where x is the profile shift coefficient (zero in the case
of polymer gears)
σF = σFO KA KV KH α KH β (9)
Ft Base helix angle
σFO = YF Ys Yβ (10)
bmn
βb = sin−1 (sin β cos αn )
A2.2.1 YF : tooth form factor z
zn = 2
cos βb cos β
The tooth form factor accounts for tooth shape
(equation (11)). Virtual number of teeth of a helical gear
 
(6hFe /mn ) cos αFen 2G π E π
YF = (11) H= − −
(SFn /mn )2 cos αn zn 2 mn 3

Proc. IMechE Vol. 224 Part L: J. Materials: Design and Applications JMDA315
Kinematic analysis of meshing polymer gear teeth 115

Hence (b) the conversion of the calculated bending stress


into a true bending stress, accounting for the
2G
θ= tan θ − H (13) exact location of the applied load, at the point of
zn maximum loading
YS = (1.2 + 0.13L)qs[1/1.21+(2.3/L)] (14)
A2.2.2 YS : stress correction factor
where
The stress correction factor (equation (14)) converts SFn SFn
the nominal bending stress to local tooth root stress L= , qs =
hFe 2ρF
and accounts for the following:
(a) the stress amplifying effect of the section change
ρf ρFp 2G 2
= +
at the tooth root fillet radius; mn mn cos θ(zn cos2 θ − 2G)

JMDA315 Proc. IMechE Vol. 224 Part L: J. Materials: Design and Applications

View publication stats

Vous aimerez peut-être aussi