You are on page 1of 19

Mechanism and Machine Theory 155 (2021) 104055

Contents lists available at ScienceDirect

Mechanism and Machine Theory


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

A study of effects of tooth surface wear on time-varying


mesh stiffness of external spur gear considering wear
evolution process
Wei Chen, Yulong Lei, Yao Fu∗, Liguo Hou
State Key Laboratory of Automotive Simulation and Control, Jilin University, Changchun, Jilin, 130025, China

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

Article history: Gear meshing surfaces are readily exposed to wear due to inadequate lubrication and long-
Received 23 February 2020 time operation. Accumulative wear causes variations of gear tooth profile, resulting in vari-
Revised 6 August 2020
ations of gear tooth’s mesh stiffness and gear system’s dynamic characteristics. This study
Accepted 6 August 2020
presents an evaluation of influences of tooth surface wear (TSW) on stiffness of the gear
mesh considering wear evolution process. Specifically, a wear prediction model for exter-
Keywords: nal spur gear (ESG) was established according to the Archard’s wear equation. Considering
Tooth surface wear the worn tooth profile, a modified analytical time-varying mesh stiffness (TVMS) model
Wear evolution process was derived by the potential energy method. The modified TVMS model was then applied
Time-varying mesh stiffness for an ESG pair and the numerical results demonstrate that TSW may lead to significant
Potential energy method reductions of mesh stiffness during the double tooth contact period. Furthermore, effects
External spur gears
of TSW on TVMS were quantified and the results indicate that the mean decrease of mesh
stiffness is almost proportional to the maximum tooth wear in the initial wear evolution
process.
© 2020 Elsevier Ltd. All rights reserved.

Abbreviations

DTP double-tooth-pair
ESG external spur gear
FEM finite element methods
GBC gear base circle
GRC gear root circle
LPTC lowest point of tooth contact
LSF load sharing factor
STP single-tooth-pair
TSW tooth surface wear
TVMS Time-varying mesh stiffness


Corresponding author.
E-mail address: fu_yao@jlu.edu.cn (Y. Fu).

https://doi.org/10.1016/j.mechmachtheory.2020.104055
0094-114X/© 2020 Elsevier Ltd. All rights reserved.
2 W. Chen, Y. Lei and Y. Fu et al. / Mechanism and Machine Theory 155 (2021) 104055

Nomenclature

Ax ,Ax the area of tooth section of perfect and worn gear


aH half contact width
b the tooth width
d0i ( i = 1, 2 ) diameter of the pitch circle, subscript i denotes the pinion and gear
d ,d  the tooth effective length of perfect and worn gear
E,Ee Young’s modulus and the equivalent Young’s modulus
E˜h wear depth function
F the total meshing force
Fa ,Fb two orthogonal decomposed forces of the total meshing force
F i ( i = 1, 2 ) meshing force of the ith tooth pair
FP the normal applied load
G shear modulus
Gw the dimensionless lubricant pressure-viscosity coefficient
h,h the distance from the mating point on the perfect and worn teeth to the tooth central line
hmin the minimum film thickness
hw wear depth
p
hwi (i = 1, 2; p = 1, 2) tooth surface wear depth along the load line, superscripts 1 and 2 denote tooth pairs 1 and 2,
and subscripts 1 and 2 refer to the pinion and gear
hx ,hx the height of the section of the perfect and worn teeth
Ix ,Ix area moment of inertia section of the perfect and worn teeth
K the time-varying mesh stiffness
ka ,kb ,kh ,k f ,ks axial compressive, bending, Hertzian contact, fillet foundation and shear stiffness of perfect
tooth
k1 ,k2 mesh stiffness of the first tooth pair and second tooth pair
ka ,kb ,ks axial compressive, bending and shear stiffness of worn tooth
kw the dimensionless wear coefficient
kw0 the wear coefficient in the boundary lubrication regime
ks ,kd the mesh stiffness during the STP and DTP engagement
LSF i (i = 1, 2) the load sharing factor of gear tooth, superscripts 1 and 2 denote tooth pairs 1 and 2
Lw the dimensionless load
N the total number of gear revolutions
n number of gear revolutions
Pb base pitch
pP the local contact pressure
pP,(n ) the local contact pressure at the same point within the nth meshing cycle
rb ,rr the radii of base circle and root circle
Rrms the composite surface roughness
Sw the dimensionless composite surface roughness RMS amplitude
sPi (i = 1, 2) the relative sliding distance, subscript i denotes mating points P 1 and P 2
T torque applied to the pinion
Uai ,Ubi ,Usi (i = p, g) axial compressive energy, bending energy, and shear energy, subscript i denotes the pinion and
gear
Uh Hertzian contact energy
Ui (i = 1, 2) the peripheral velocity, subscript i denotes mating points P 1 and P 2
Utotal total potential energy of a single pair of the meshing teeth
x,x the distance from the tooth section to the tooth root of perfect and worn gear
y the distance from the meshing point to the pitch point
yi the distance from the Hertz contact center

Greek symbols
α0 the pressure angle of the pitch circle
α1 the angle between the action force F and the decomposed force Fb
α2 the minus half of the tooth angle measured on GBC
α3 , α4 minus half angles measured on the GRC in Scenario 1 and Scenario 2
α5 the angle between the action force F and the decomposed force Fb when the meshing point is on the
GRC
αsi (i = 1, 2) minus half angle measured on the LPTC, subscript i denotes Scenario 1 and Scenario 2
W. Chen, Y. Lei and Y. Fu et al. / Mechanism and Machine Theory 155 (2021) 104055 3

α angular displacement of arbitrary point at the involute curve


β the pressure angle of the LPTC
δ i ( i = 1, 2 ) deformation of the ith tooth pair
θ angular displacement of the pinion or gear
ν Poisson’s ratio
λ the ratio of minimum film thickness to composite surface roughness
φ the evolving angle of the LPTC
ωi ( i = 1, 2 ) the angular velocity, subscript i denotes the pinion and gear

1. Introduction

Gearboxes are widely employed for power transmission in wind turbines, helicopters and automobiles [1]. As a core
component of gearbox, gears may suffer from failures due to heavy load, harsh operating conditions or fatigue [2]. Indeed,
the severity of gear defect determines the performance of gearbox. Hence, the state of gears shall be evaluated to guarantee
normal operation of gearbox [3]. Closely related to the state of gears [4], time-varying mesh stiffness (TVMS) can be utilized
to derive characteristics of dynamic response [5] and the state of gears can be assessed by the features of TVMS [3].
TVMS calculation has attracted extensive attention over the past decades. According to previous studies, TVMS calculation
methods can be categorized as empirical formula methods [6,7], analytical methods [8–14], finite element methods (FEM)
[15–18], analytical-FEM [19–21], and experimental methods [22–25]. Among these methods, empirical formula methods are
suitable for rapid yet low-precision calculations of mesh stiffness [26], while reliable and accurate experimental methods
require specific measurement approaches and equipment [27]. The other three methods are widely employed for evaluation
of TVMS of gears with faults such as tooth crack, pitting, spalling and tooth wear. When gear tooth surface wear (TSW)
causes irregular tooth profile errors, FEM is complicated and time-consuming [3]. The analytical method is a facile yet
effective one with reasonable accuracy [28]. Yang et al. [8] calculated the mesh stiffness based on the Timoshenko beam
theory and the potential energy principle using an analytical method. In this method, the gear tooth was denoted by a
cantilever beam model and it was assumed that the total energy is stored in a meshing tooth as axial compressive energy,
bending energy and Hertzian energy. This model was further modified by considering the shear energy [9]. To enhance the
accuracy of TVMS calculation, Chen et al. [10] reported a method combining the fillet foundation deflection formula and the
potential energy [29]. Then, an analytical model was established for calculation of mesh stiffness considering tooth profile
and spacing errors [11]. In this study, an analytical model is developed to evaluate TVMS considering TSW.
As the most commonly observed failure on working surface of gear tooth [30], wear has attracted great attention over
the past decades. A mathematical wear model, which is known as Archard’s wear equation, was reported by Archard [31].
The Archard’s wear equation has been widely employed for calculations of tooth flank surface wear in gear mechanisms
[32–37]. Flodin et al. [32,33] developed an analytical model for prediction of tooth wear in spur gears. Subsequently, wear
calculation models were developed for helical gears and gear teeth were regarded as thin uncoupled slices [34,35]. Bajpai et
al. [36] reported a method for wear prediction of contacting tooth surface by combining a FE-based gear contact model and
the Archard’s wear formula. Ding et al. [37] proposed a model to predict gear teeth wear considering effects of lubrication
conditions on the wear coefficient. Brandão et al. [38] reported wear tests aimed at investigation of effects of load and lubri-
cation on the wear coefficient. As demonstrated by studies mentioned above, the calculation results by the Archard’s wear
equation are mainly dependent on load distribution and lubrication condition. To improve the accuracy of wear prediction
model, the cycle-varying load distribution along the tooth profile and the lubrication state are considered in this study.
In recent years, great efforts have been devoted to investigations of effects of tooth wear on TVMS [3,39–41]. Feng et al.
[39] developed a computational method for TVMS considering TSW and investigated the effects of wear severity on TVMS.
Shen et al. [40] reported an analytical method for calculation of TVMS of an external spur gear (ESG) pair considering TSW.
Similarly, the effects of TSW on TVMS of a planetary gear set were evaluated [3]. Sun et al. [41] established a calculation
model for TVMS of spur gears and investigated the effects of tooth flank wear on mesh stiffness. In aforementioned studies,
the mesh stiffness during the double-tooth-pair (DTP) meshing period is obtained by accumulating the stiffness of all gear
pairs in mesh even if in presence of tooth profile errors caused by wear. Additionally, TVMS of gear with TSW considering
the wear evolution process has not been thoroughly studied. Indeed, wear is not an instantaneous catastrophic fault; instead,
it undergoes a long-term evolution process [42], during which gear tooth wear and TVMS are highly dependent on each
other. The gear TSW contributes to the variation of tooth profile and alters the TVMS, which in turn affects TSW by altering
distribution of tooth flank load [43].
Aiming at the interaction between TVMS and TSW, a modified TVMS model for spur gear considering the wear evolu-
tion process is proposed. The potential energy method was utilized to develop the mesh stiffness formula for a single ESG
tooth and the total mesh stiffness was calculated with TSW taken into consideration. Based on the TVMS model, a load
sharing factor (LSF) model considering TSW was obtained. Finally, the wear depth was calculated by substituting the load
distribution into Archard’s wear equation, in which the boundary lubrication condition is also taken into consideration.
4 W. Chen, Y. Lei and Y. Fu et al. / Mechanism and Machine Theory 155 (2021) 104055

In this article, Section 2 presents an analytical model to evaluate gear TSW based on the Archard’s wear equation.
Section 3 presents a modified TVMS of ESG with TSW taken into consideration derived using the potential energy method.
Section 4 presents a case study that demonstrates the effects of TSW on TVMS of ESG. Section 5 draws a conclusion.

2. Wear prediction model of ESGs

In this section, the Archard’s wear equation, a widely recognized wear prediction method, is employed for calculation of
TSW depth of ESGs. The general Archard’s wear model of a local contact point (P) on meshing surfaces can be expressed as
follows [30]:
 sP
hw P = kw p p d sP (2.1)
0

where hwP represents the wear depth, kw represents the dimensionless wear coefficient, sp represents the sliding distance,
and pp represents the local contact pressure.
According to the concept of single point observation [32], the additional wear at point P of the meshing surface after the
nth meshing cycle may be described as:


hwP ,(n) = kw pP,(n) sP (2.2)

where
hw P ,(n ) denotes the additional TSW depth at Point P after the nth meshing cycle, sP represents the sliding distance
of Point P, and pP, (n) gives the local contact pressure at the same point within the nth meshing cycle.
The wear coefficient is determined by the load, lubrication condition, surface roughness, and material properties, as in
[37]:

kw0 , λ < 0.5 (Boundary lubrication )
kw = 2kw0 (4 − λ )/7, 0.5 < λ < 4 (Mixed lubrication ) (2.3)
0, λ > 4 (Elastohydrodynamic lubrication )
where λ is the ratio of minimum film thickness (hmin ) to composite surface roughness (Rrms ), and kw0 is the wear coefficient
in the boundary lubrication regime at relatively low speed [44]:

kw0 = 3.981 × 1029 L1w.219 G−7


w
.377 1.589
Sw /Ee (2.4)

where Ee refers to the equivalent Young’s modulus. Lw , Sw and Gw represent the dimensionless load, the dimensionless
composite surface roughness RMS amplitude, and the dimensionless lubricant pressure-viscosity coefficient, respectively.
The detailed formulae are given elsewhere [44].
The relative sliding distance for mating points P1 on pinion and P2 on gear can be written respectively as [30]:
U − U 
1 2
sP1 = 2aH (2.5)
U1
U − U 
2 1
sP2 = 2aH (2.6)
U2

where aH represents the half contact width given in Appendix A, U1 and U2 denote peripheral velocities of the two mating
points, respectively, which can be expressed as [30]:
 
d01
U1 = ω1 sin α0 + y (2.7)
2
 
d02
U2 = ω2 sin α0 − y (2.8)
2

where ω denotes the angular velocity, d0 and α 0 represent the diameter and the pressure angle of the pitch circle, re-
spectively, y is the distance from the meshing point to the pitch point, and subscripts 1 and 2 refer to pinion and gear,
respectively. Based on the Hertzian contact theory, the local contact pressure (pP ) along the whole sliding distance can be
given as [3]:
2FP  2 1 / 2
pP = aH − y2i (2.9)
π baH
2

where FP is the normal applied load, b denotes the active gear tooth width, and yi denotes the distance from the Hertz
contact center. In this study, the Hertz contact center is employed as the contact point. According to Eq. (2.9), pP is propor-
tional to FP . However, the variation of tooth profile induced by TSW can lead to changes of the load distribution along tooth
W. Chen, Y. Lei and Y. Fu et al. / Mechanism and Machine Theory 155 (2021) 104055 5

Fig. 1. Tooth surface engagement region of an ESG.

profile [11]. Hence, a LSF model is used to describe the change of load, as will be discussed in Section 3. Therefore, FP can
be calculated by:

FP = LSF · F (2.10)

T
F= (2.11)
rb
where F refers to the total meshing force whose direction is tangent to the gear base circle (GBC), T demonstrates the
applied torque, and rb denotes the radius of the GBC.

3. Derivation of mesh stiffness for ESGs considering TSW

For external gear pairs with contact ratio of 1-2, the tooth surface can be divided into the double-tooth engagement
region and the single-tooth engagement region (see Fig. 1). Correspondingly, the mesh stiffness in the tooth engagement
includes two terms: mesh stiffness during the single-tooth-pair (STP) period and mesh stiffness during the DTP period [9].
To obtain the mesh stiffness during the entire meshing period, Sections 3.1 and 3.2 depict the mesh stiffness during the STP
engagement and the mesh stiffness during the DTP engagement, respectively.

3.1. Mesh stiffness in the STP engagement

In the STP engagement, only one pair of teeth are engaging. A single tooth is modeled as a non-uniform cantilever beam
and the mesh stiffness is calculated by the potential energy method [2]. The tooth energy stored in a STP can be calculated
by [27]:

p g p g F2
Utotal = Uh + Utooth + Utooth + Ubody + Ubody = (3.1)
2k
i
where Utooth (i = p, g) represents the energy in case the gear tooth is regarded as a cantilever beam with varying cross-
i
section, Uh refers to the Hertzian contact energy, Ubody (i = p, g) denotes the energy of the elastic gear body, F dedicates the
total meshing force, k represents the mesh stiffness, and superscripts p and g denote pinion and gear, respectively.
According to the Timoshenko beam theory [45], Utoothi (i = p, g) can be expressed as:
i
Utooth = Ubi + Usi + Uai (3.2)
where Ubi , Usi , and Uai represent the bending energy, shear energy, and axial compressive energy stored in a single tooth,
respectively.
As shown in Fig. 2, TSW only leads to variation of tooth profile [3]. Hence, only the energy stored in the tooth varied.
TSW alters all three mesh stiffness terms. Considering TSW, the new analytical expressions for the three mesh stiffness
terms are derived below.

3.1.1. Three mesh stiffness terms of ESGs considering TSW


Previously, the potential energy method has been employed to derive the energy stored in a single gear tooth [2,14,27,46].
In these models, the ESG tooth is modeled as a cantilever beam fixed on the gear root circle (GRC), as shown in Fig. 3. The
three terms of energy stored in a perfect gear tooth with meshing force can be generally described as:

[F (d − x ) cos α1 − F h sin α1 ]
d 2
F2
Ub = = dx (3.3)
2kb 0 2E Ix
6 W. Chen, Y. Lei and Y. Fu et al. / Mechanism and Machine Theory 155 (2021) 104055

Fig. 2. Variation of the tooth profile caused by TSW [3].

Fig. 3. Model of a gear tooth without TSW.


F2 d
3F 2 cos2 α1
Us = = dx (3.4)
2ks 0 5GAx


F 2 sin α1
d 2
F2
Ua = = dx (3.5)
2ka 0 2E Ax
where E refers to the Young’s modulus, G refers to the shear modulus, kb , ks and ka denote the bending, shear, and axial
compressive mesh stiffness, respectively, h refers to the distance from the mating point on gear to the tooth central line, d
indicates the distance from the contact point on gear to the gear root, Ax is the area of the tooth section, Ix refers to the
moment of inertia of the tooth section, and x is the distance from the tooth section to the tooth root.
According to previous studies [3,14,28], two scenarios shall be considered in calculation of the mesh stiffness of a stan-
dard external spur gear with the addendum coefficient of 1, tip clearance coefficient of 0.5 and pressure angle of 20◦ .

Scenario 1: GRC radius < GBC radius.

When less than 42 teeth are present, GRC radius is smaller than GBC radius (rr < rb , see Fig. 4). The total energy stored
in the tooth includes two parts: the energy stored in the part between contact section and GBC and the energy stored in
the part between GRC and GBC.
In Fig. 4, the dotted line along the involute profile represents the worn tooth profile, which varies from that of the
perfect involute with the wear depths along the load line. As the contact points cannot reach the lowest point of involute
profile and there is no wear on the tooth profile between the lowest point of the involute profile and the lowest point of
W. Chen, Y. Lei and Y. Fu et al. / Mechanism and Machine Theory 155 (2021) 104055 7

Fig. 4. Model of a gear tooth considering TSW in Scenario 1.

tooth contact (LPTC), the energy stored in the part between contact section and GBC can be further divided into two parts
considering TSW. Therefore, the three terms of mesh stiffness can be calculated with three components, respectively:

2
1 d1 (d − x ) cos α1 − h sin α1
= dx
k b 0 E Ix

2
2
 ds (d − x ) cos α1 − h sin α1  d (d − x ) cos α1 − h sin α1
+ dx + dx (3.6)
d1 E Ix ds E I x

   d

1 d1
6cos2 α1 ds
6cos2 α1 6cos2 α1 
= dx + dx + dx (3.7)
k s 0 5GAx d1 5GAx ds 5GA x
   d

sin α1 sin α1 sin α1 
d1 2 ds 2 2
1
= dx + dx + dx (3.8)
k a 0 E Ax d1 E Ax ds E A x

where symbol ‘ ’ means that the variables consider the effects of TSW and ds represents the distance from the LPTC to the
gear root. In Eqs 3.6–3.8, x , d , hx , and h are shown in Fig. 4 and have been derived as in Appendix B (Note: α 2 is defined
as a negative value in this study):

x = rb [(α − α2 ) sin α + cos α ] − rr cos α3 − hwx sin α (3.9)

d = rb [(α1 − α2 ) sin α1 + cos α1 ] − rr cos α3 − hw sin α1 (3.10)

hx = rb [(α − α2 ) cos α − sin α ] − hwx cos α (3.11)

h = rb [(α1 − α2 ) cos α1 − sin α1 ] − hw cos α1 (3.12)

where rb and rr denote GBC radius and GRC radius, respectively, α is the pressure angle of reference circle, hw and hwx
respectively represent the wear depths of tooth contact section and the tooth section where the distance from the gear
tooth root is x. It should be noted that hw and hwx varied with the section position.
By differentiating Eq. (3.9), dx can be calculated by:

dx = [rb (α − α2 ) cos α − ( (dhwx /dα ) sin α + hwx cos α )]dα (3.13)
Due to the presence of single-sided wear, the moment of inertia and the cross-section area can be determined by:
1  3
Ix = hx + h x b (3.14)
12

Ax = hx + h x b (3.15)
8 W. Chen, Y. Lei and Y. Fu et al. / Mechanism and Machine Theory 155 (2021) 104055

Fig. 5. Model of a gear tooth considering TSW in Scenario 2.

where b refers to the tooth width, and hx denotes the cross-section height of gear without wear, which can be determined
by:

|rb sin α2 |, if 0 ≤ x ≤ d1
hx = (3.16)
rb [(α − α2 ) cos α − sin α ], if d1 ≤ x ≤ d
To evaluate the effects of TSW on mesh stiffness, the relationship may be expressed as angular variables for the profile
with wear. Substituting Eqs. (3.9) to (3.16) into Eqs. (3.6) to (3.8), the three terms of mesh stiffness of spur gear with TSW
can be determined by:

2  αs1
1 d1 3 (d − x ) cos α1 − h sin α1 3D2 (α − α2 ) cos α
= dx + dα
k b 0 2Ebrb3 |sin α2 |
3
α2 2Eb[(α − α2 ) cos α − sin α ]
3

 α1
12(Drb + hwx cos α1 sin α )
2
+ [rb (α − α2 ) cos α − S]dα (3.17)
αs1 EbH 3

  αs1
1 6(1 + ν )cos2 α1
d1
6(1 + ν )(α − α2 ) cos α cos2 α1
= dx + dα
k s 0 5 Ebr b |sin α2 | α2 5Eb[(α − α2 ) cos α − sin α ]
 α1
12(1 + ν )cos2 α1
+ [rb (α − α2 ) cos α − S]dα (3.18)
αs1 5EbH

  αs1  α1
sin α1 (α − α2 ) cos α sin2 α1 sin α1
d1 2 2
1
= dx + dα + [rb (α − α2 ) cos α − S]dα (3.19)
k a 0 2Ebrb |sin α2 | α2 2Eb[(α − α2 ) cos α − sin α ] αs1 EbH
where D = 1 + cos α1 [(α2 − α ) sin α − cos α ], S = (dhwx /dα ) sin α + hwx cos α , H = 2rb [(α − α2 ) cos α − sin α ] − hwx cos α , ν
denotes Poisson’s ratio, and α s1 represents the minus half angle measured on the LPTC in Scenario 1 and can be expressed
as follow:
αs1 = α2 + φ (3.20)
where φ is the evolving angle of the LPTC, which has been given in Appendix B.

Scenario 2: GRC radius > GBC radius.

If more than 42 teeth are present, GRC radius is larger than GBC radius (rr > rb , see Fig. 5). The total energy stored in
the tooth equals to the energy stored in the part between the contact section and the root section.
Correspondingly, the three terms of mesh stiffness of gear tooth considering the influence of wear can be written as:

2 
2
1 ds (d − x ) cos α1 − h sin α1 d (d − x ) cos α1 − h sin α1
= dx + dx (3.21)
k b 0 E Ix ds E I x
W. Chen, Y. Lei and Y. Fu et al. / Mechanism and Machine Theory 155 (2021) 104055 9

  d
1 ds
6cos2 α1 6cos2 α1 
= dx + dx (3.22)
k s 0 5GAx ds 5GA x
  d
sin α1 sin α1 
ds 2 2
1
= dx + dx (3.23)
k a 0 E Ax ds E A x
where symbol ‘ ’ means that the variables consider the wear depth. The distance from the arbitrary point at the involute
curve to GRC (x ), and the distance from the mating section to GRC (d ) are determined by:
x = rb [(α − α2 ) sin α + cos α ] − rr cos α4 − hwx sin α (3.24)

d = rb [(α1 − α2 ) sin α1 + cos α1 ] − rr cos α4 − hw sin α1 (3.25)


where α 4 represents the minus half angle measured on the GRC in Scenario 2 and can be expressed as follow:
  r 
α4 = inv arccos b + α2 (3.26)
rr
Therefore, the three terms of mesh stiffness can be described as follows, respectively:
 αs2  α1
3D2 (α − α2 ) cos α 12(Drb + hwx cos α1 sin α )
2
1
= d α + [rb (α − α2 ) cos α − S]dα (3.27)
k b α5 2Eb[(α − α2 ) cos α − sin α ]
3
αs2 EbH 3
 αs2  α1
1 6(1 + ν )(α − α2 ) cos α cos2 α1 12(1 + ν )cos2 α1
= dα + [rb (α − α2 ) cos α − S]dα (3.28)

ks α5 5Eb[(α − α2 ) cos α − sin α ] αs2 5EbH

 αs2
1 (α − α2 ) cos α sin2 α1
= dα
α5 2Eb[(α − α2 ) cos α − sin α ]

ka
 α1
sin α1
2
= [rb (α − α2 ) cos α − S]dα (3.29)
αs2 EbH
where α 5 is the angle between the action force F and the decomposed force Fb when the meshing point is on the GRC, and
α s2 represents the minus half angle measured on the LPTC in Scenario 2. According to the geometric relationships in Fig. 5,
α 5 and α s2 can be expressed as follow, respectively:
r 
α5 = arccos b
+ α4 (3.30)
rr

αs2 = α2 + φ + β (3.31)
where φ and β represent the evolving angle and pressure angle of the LPTC, which have been given in Appendix B, respec-
tively.

3.1.2. Hertzian contact and fillet foundation stiffness


According to previous studies [47], the Hertzian contact stiffness of two meshing teeth is regarded as a constant along
the load line and it is dependent on neither the contact position nor the interpenetration depth:

1 4 1 − ν2
= (3.32)
kh π Eb
The fillet-foundation deflection also affects the mesh stiffness. Sainsot et al. [29] proposed an analytical formula based
on Muskhelishvili’s elastic ring theory to calculate the stiffness of the flexible base:
  2   
1 cos2 αm ∗ μf μf 
= L +M ∗ ∗
+ P 1 + Q tan ∗ 2
αm (3.33)
kf Eb Sf Sf

The details of parameters are given in a previous study [29].

3.1.3. Total effective mesh stiffness in the STP engagement


The three terms of mesh stiffness of spur gear considering TSW have been derived in Section 3.1.1, while the Hertzian
contact stiffness and the fillet foundation stiffness are presented in Section 3.1.2. Then, the total effective mesh stiffness
considering TSW in the STP period can be obtained by:
1
k= 1 1 1 1 1 1 1 1 1
(3.34)
kh
+ k
+ k
+ k
+ kf1
+ k b2
+ k  s2
+ k a2
+ kf2
b1 s1 a1
10 W. Chen, Y. Lei and Y. Fu et al. / Mechanism and Machine Theory 155 (2021) 104055

Fig. 6. Schematic illustration of two tooth pairs along the action line with TSW.

3.2. Mesh stiffness in the DTP engagement

As mentioned in Section 2, wear causes the tooth profile error compared with the perfect involute tooth. The engagement
of wore gear teeth is not consistent with that of ideal involute gear teeth in the double-tooth period. Hence, the mesh
stiffness during DTP period cannot be obtained via summation of meshing stiffness components during STP period [11].
Based on a previously reported model [11], models for DTP mesh stiffness and LSF considering TSW are described briefly in
this section.
The DTP engagement along the action line with consideration of TSW is displayed in Fig. 6. Herein, Pb denotes the ideal
base pitch. hw and δ represent TSW depth along the load line and the tooth pair deformation, respectively. Superscripts 1
and 2 denote tooth pairs 1 and 2, respectively. Subscripts 1 and 2 refer to the pinion and gear, respectively. In the gear pair,
the gear is stationary while the pinion rotates freely. If the pinion is under load, it will rotate to engage and tooth pairs 1
and 2 will deform (see the red region in Fig. 6). If two gear pairs are simultaneously in mesh, the deformation relationship
between the gear pairs and meshing force (F) can be expressed as:

δ 1 − δ 2 = E˜h = h2w2 + h2w1 − h1w2 − h1w1 (3.35)

F 1 = k1 δ 1 , F 2 = k2 δ 2 (3.36)

F = F1 + F2 (3.37)

where hw is positive as material is removed from the contact point of gear surface due to wear, E˜h is the wear depth
function considering the effect of all wear of the gear pairs in mesh, ki (i = 1, 2) is a function of tooth contact deformation
and angular displacement of the pinion or gear (θ ), which is given in Eq. (3.38) and δ i = 0 indicates that the teeth are
independent of each other. Hence, the effective mesh stiffness is zero.

 ki θ , δ i > 0
ki =ki θ , δi = ( ) i (3.38)
0, δ =0
Hence, the mesh stiffness during the DTP engagement can be expressed as [11]:
 k1 +k2
1+k2 E˜h /F
, δ1 > δ2
k = F /δ
d d
max = (3.39)
k1 +k2
1−k1 E˜h /F
, δ1 < δ2

LSF, which is defined as the ratio of the meshing force of a tooth pair to the total meshing force, can be determined by
[11]:
 
1F1 k1 k2 E˜h
LSF = = 1 1+ (3.40)
F k + k2 F
 
2F2 k2 k1 E˜h
LSF = = 1 1− (3.41)
F k + k2 F
W. Chen, Y. Lei and Y. Fu et al. / Mechanism and Machine Theory 155 (2021) 104055 11

Fig. 7. Flow chart of calculation of TVMS model considering TSW.

3.3. Total mesh stiffness of spur gears considering TSW

The flow chart for the TVMS of spur gear is illustrated in Fig. 7. It combines the TSW depth model and the TVMS model.
When the wear of tooth surface reaches a threshold, a phase change of the mesh stiffness may be observed, but it is not
taken into consideration. Therefore, the mesh stiffness in the entire meshing period can be given by:

ks , single − tooth engagement duration
K= (3.42)
kd , double − tooth engagement duration

4. Case study

According to previous studies [10,11,30], Sections 2 and 3 present analytical calculation models of TSW and TVMS of an
ESG, respectively. In this section, a case study is involved to investigate the effects of tooth wear on the gear mesh stiffness.
Specifically, wear analysis and validation are conducted, followed by quantification of the effects of tooth wear on the gear
mesh stiffness. Table 1 summarizes gear pair parameters and simulation conditions.

4.1. Wear results analysis and verification using experiment data

As mentioned in Section 2, the wear depth is mainly affected by the contact load and wear coefficient under constant
operating conditions. Fig. 8 illustrates LSF of the gear with different number of gear revolutions (N). As observed, LSF of the
gear pair changes with the gear revolution number, indicating that LSF of the gear pair varies with the cumulative wear
depth. As the TSW depth increases, the variation of contact load between teeth increases. This is consistent with previous
studies [43]. Additionally, the wear coefficient is crucial to calculation of wear depth, which is assumed to be a constant in
most studies. Based on Eq. (2.4), it is assumed in this study that the wear coefficient varies with the meshing position and
load, as shown in Fig. 9. In STP period, the wear coefficient increases as the gear tooth bears the overall load. After several
revolutions, the wear coefficient changes due to the variation of LSF caused by TSW (see Fig. 8).
Fig. 10 shows the cumulative wear depth of all meshing positions on the pinion and gear with different N values. As
shown in Fig. 10(a), severe wear was observed in the pinion tooth root region due to large wear coefficient, large contact
pressure and high sliding ratio. As a result, the wear depths over the teeth surfaces varied with the maximum wear at
the gear tooth tip (see Fig. 10(b)). Wear was not observed on the pitch line of gear pair since relative slide was absent at
12 W. Chen, Y. Lei and Y. Fu et al. / Mechanism and Machine Theory 155 (2021) 104055

Table 1
Parameters in the simulations [43].

Items Pinion/ Gear

Tooth number 16/24


Module (mm) 4.5
Tooth width (mm) 14
Pressure angle (deg) 20
Addendum coefficient 1
Tip clearance coefficient 0.5
Theoretic contact ratio 1.55
Poisson’s ratio 0.3
Young’s modulus (Pa) 2.1 × 1011
Inner bore radius (mm) 20/20
Transmitted torque (Nm) 302
Rotation speed of the pinion (rpm) 100
Composite surface roughness (μm) 0.3
Pressure-viscosity coefficient of the lubrication oil (m2 /N) 2.6 × 10−8 [48]

Fig. 8. LSF of the gear pair.

Fig. 9. Wear coefficient of the gear pair (kw ).

this position. Notably, TSW depths of the gear pair changed suddenly at moment A due to the alternation of single teeth
engagement and double teeth engagement. These conclusions are consistent with previous studies [43,49].
The amount of wear at the tooth root of the pinion with different running times are depicted in Fig. 11. The variation
rule of wear depth from the simulation is similar to that obtained by the experiment in Ref. [49] (see Fig. 11), which further
verifies the proposed method. Two adjacent teeth are measured to obtain the experiment data and the differences between
the two measured results may be caused by manufacture and assembly errors. But these errors are ignored and the teeth
are assumed to be identical in the simulation. The wear rate slows down with the wear evolution, and this phenomenon
is more obvious for the measured results. No additives in the lubrication oil may lead to a higher initial wear rate in the
experiment.
W. Chen, Y. Lei and Y. Fu et al. / Mechanism and Machine Theory 155 (2021) 104055 13

Fig. 10. TSW depth of the gear pair under different revolutions.

Fig. 11. The amount of wear at the root of the pinion.

4.2. TVMS of external spur gears considering TSW

In this subsection, the mesh stiffness model established in this study is validated by FEM results from Ref. [43] and the
effects of TSW on TVMS are quantified.

4.2.1. Validation of the mesh stiffness model considering TSW


The mesh stiffness of the gear pair considering TSW can be calculated using equations in Section 3. In Ref. [43], a two-
dimensional (2-D) finite element model was established to evaluate the mesh stiffness of gear pair with irregular worn tooth
14 W. Chen, Y. Lei and Y. Fu et al. / Mechanism and Machine Theory 155 (2021) 104055

Table 2
Mesh stiffness comparison

Maximum difference (%) Mean difference(%)


Gear wear cycle
STP period DTP period One mesh period

N=0 3.1 12.6 5.3


N = 2.25 × 105 3.1 12.2 4.5

Fig. 12. Comparison of the mesh stiffness calculated by FEM in Ref [43] and the proposed method in this study.

Fig. 13. TVMS of the gear pair considering TSW.

profiles. The gears were meshed using Plane 182 and the contact effect was modeled using Conta174 and Targe170. Table 2
and Fig. 12 give the comparisons between the mesh stiffness results by FEM and that obtained by the proposed method in
this study. Comparing the results by these two methods, the maximum difference in the STP period is 3.1% and that in the
DTP period is 12.6%, which is mainly due to the neglect of the structure coupling effect in the DTP period. The change trend
of mesh stiffness considering TSW is consistent with FEM method. The difference in amplitude change is mainly due to the
fact that only the worn mesh points are assumed to contact to obtain the stiffness in the potential energy method, while
the contact of the tooth profile around the worn mesh points is considered as well by FEM. The proposed method costs
about 1 second to complete the mesh stiffness calculation for one mesh cycle, while the FEM may cost about 2 hours [43].
Therefore, the proposed method can be used to calculate the stiffness variation in the wear evolution more efficiently and
relatively accurately.

4.2.2. Effect of TSW on TVMS of external spur gears


Fig. 13 shows the mesh stiffness of the gear pair. As observed, the TVMS dropped due to TSW. The effects of TSW on
mesh stiffness during the DTP period are more significant than those in the STP period. This is consistent with results
reported elsewhere [43]. In theory, TSW causes the variation of gear tooth profile, thus altering the gear pair engagement
and increases the maximum deformation. Thereby, the tooth stiffness calculated by Eq. (3.38) would decrease during the
W. Chen, Y. Lei and Y. Fu et al. / Mechanism and Machine Theory 155 (2021) 104055 15

Fig. 14. The mean mesh stiffness reduction and the maximum wear depth of the gear pair.

DTP period. Indeed, the mesh stiffness also decreases in the STP period, although it is not significant [3]. The reason is that
the wear depth is not comparable to geometric parameters of the gear pair in the initial wear evolution process.
As mentioned in Section 4.1, wear on the pinion was much more severe than that on the gear and the maximum wear
was observed at the pinion tooth root. Therefore, wear at the pinion tooth root was used to describe the wear severity of
tooth surface of gear pair, which will be utilized to depict the effects of TSW on TVMS.
To quantify effects of TSW on gear mesh stiffness, the mean mesh stiffness reduction corresponding to different wear
severities are calculated. Fig. 14 gives the relation between TSW depth and the mean mesh stiffness reduction. The max-
imum wear depth, the mean mesh stiffness reduction and the percentage reduction in mean stiffness are depicted in
Fig. 14(a). As observed, the slope of the wear depth gradually decreases with the wear evolution, which is mainly attributed
to the decrease of contact load at the pinion root. Additionally, the mean mesh stiffness reduction has the same changing
tendency as the wear depth. Fig. 14(b) shows that the mean mesh stiffness reduction in one mesh cycle is almost propor-
tional to the maximum wear depth in the initial wear evolution process.

5. Conclusions

This study provides an analytical model to quantify the effects of TSW on gear mesh stiffness. Herein, a wear prediction
model was incorporated into a TVMS model. The wear model was established according to the Archard’s wear equation, in
which the varying wear coefficient and load distribution are considered. Considering the wear profile, the modified analytical
TVMS model is derived using the potential energy method. The effects of TSW on the TVMS are demonstrated by a case
study. In the case study, the wear depth agrees well with the experiment data in a published literature, which validates
the accuracy of the wear depth model in this study. The stiffness results are compared with that calculated by FEM, which
indicates that the mesh stiffness model for a single tooth pair is accurate. The simulation results show that the decrease of
TVMS during the DTP period was significantly larger than that during the STP period, which is consistent with FEM results
as well. Additionally, the mean mesh stiffness reduction in one mesh cycle is almost proportional to the maximum TSW (at
the dedendum of the pinion tooth) if N ≤ 5 × 105 . The proposed model has the potential to be integrated into the gear
dynamic model, by which the dynamic influence caused by gear TSW can be estimated.
16 W. Chen, Y. Lei and Y. Fu et al. / Mechanism and Machine Theory 155 (2021) 104055

However, the proposed model needs further improvements. For instance, the fillet foundation stiffness during the DTP
period should be revised by considering the structure coupling effect, as in a previous study [27]. However, it is time-
consuming and barely possible to incorporate a customized FE model into the proposed model. Therefore, the primary
objective of this study is clarification of effects of TSW on mesh stiffness of spur gears considering the wear evolution
process, which is seldomly considered in previous studies. In future works, an effective and efficient calculation method of
the fillet foundation stiffness during the DTP period deserves more efforts. Additionally, the phase change caused by TSW
and the effect of EHL on wear shall be considered to further improve accuracy of the proposed model.

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have
appeared to influence the work reported in this paper.

Acknowledgments

This work is supported by the International S&T Cooperation Program of China (No. 2014DFA71790), the National Key
Research and Development Program of China (No. 2018YFB0104901), the Key Scientific and Technological Project of Jilin
Province (No. 20160519008JH and No. 20170204073GX), the Science and Technology Planning Project of Jilin Province
(20180520071JH), and Science and Technology Program of Qingdao (No. 18-1-2-17-zhc). We appreciate valuable comments
and suggestions provided by reviewers and editors.

Appendix A


4FP ρ
aH = (A.1)
π bE ∗

Fig. A1. Equivalent rollers of radius ρ 1 and ρ 2 at the contact point P of involute gear tooth profile.
W. Chen, Y. Lei and Y. Fu et al. / Mechanism and Machine Theory 155 (2021) 104055 17

where
1 1 1
= + (A.2)
ρ ρ1 ρ2
d01 d02
ρ1 = sin α0 + y, ρ2 = sin α0 − y (A.3)
2 2

y = rb1 (tan α p − tan α0 ) (A.4)

1 1 − ν12 1 − ν22

= + (A.5)
E E1 E2

Appendix B

A coordinate system (XOY) is located at the gear center, shown as Fig. B1. The X-axis is along the teeth central line and
the Y-axis is vertical to the X-axis. Point E refers to the intersection of the involute curve and GBC. Indeed, the tooth profile
between GRC and GBC is not an involute curve. Hence, it can hardly be analytically described. For this reason, straight lines
AE and BF are employed to simplify the tooth profile [2]. Point D is a point on the involute curve and DC is tangent to the
GBC at point C. D is the intersection of DC and the wore tooth profile. Point P is the contact point of two mating gear teeth
and PQ is tangent to GBC at point Q. P is the intersection of PQ and the wore tooth profile. Point S is the LPTC and ST is
tangent to the GBC at point T.
In the XOY, the coordinates of points A and C can be expressed as:
 
xA = rr cos α3 xC = rb cos α
, (B.1)
yA = rr sin α3 yC = rb sin α

where rr and rb denote GRC radius and GBC radius, respectively, α is the angle between OC and X-axis (α is not negative
when C is in the quadrant I, otherwise, α is negative.), α 2 and α 3 are the minus half of the tooth angle measured on GBC
and the minus approximated half of the tooth angle measured on GRC, respectively. α 2 and α 3 can be expressed as follows:
π 
α2 = − + tan α0 − α0 (B.2)
2Z
 
rb sin α2
α3 = arcsin (B.3)
rr

where Z refers to the quantity of spur gear teeth and α 0 refers to the pressure angle.
Based on features of the involute curve, the length of DC can be obtained by:

DC = CE = rb (α − α2 ) (B.4)

Fig. B1. Model of a gear tooth considering TSW.


18 W. Chen, Y. Lei and Y. Fu et al. / Mechanism and Machine Theory 155 (2021) 104055

The coordinates of points D and D can be expressed as:


 
xD = rb cos α + rb (α − α2 ) sin α xD = rb cos α + [rb (α − α2 ) − hwx ] sin α
, (B.5)
yD = rb sin α − rb (α − α2 ) cos α yD = rb sin α − [rb (α − α2 ) − hwx ] cos α
where hwx is the wear depth along the line DC and it can be described as a function of the angle α . Then, variables x and
hx can be expressed as:
x = xD − xA = rb cos α + [rb (α − α2 ) − hwx ] sin α − rr cos α3 (B.6)

hx = −yD = [rb (α − α2 ) − hwx ] cos α − rb sin α (B.7)


Similarly, d and h can be obtained by:
d = rb cos α1 + [rb (α1 − α2 ) − hw ] sin α1 − rr cos α3 (B.8)

h = [rb (α1 − α2 ) − hw ] cos α1 − rb sin α1 (B.9)


where α 1 is the angle between OQ and X-axis (α 1 is positive when Q is in the quadrant I, otherwise, α 1 is negative.), hw
and hwx are the wear depth along the lines PQ and DC, respectively.
In Fig. B1, φ and β represent the evolving angle and pressure angle of point S, respectively. According to features of the
involute curve, φ can be expressed as:
φ = tan β − β (B.10)
where β can be calculated by the Eq. (2.45) in Ref. [9].

Appendix C

References

[1] A. Saxena, M. Chouksey, A. Parey, Measurement of FRFs of coupled geared rotor system and the development of an accurate finite element model,
Mech. Mach. Theory 123 (2018) 66–75, doi:10.1016/j.mechmachtheory.2018.01.010.
[2] X. Liang, H. Zhang, L. Liu, M.J. Zuo, The influence of tooth pitting on the mesh stiffness of a pair of external spur gears, Mech. Mach. Theory 106 (2016)
1–15, doi:10.1016/j.mechmachtheory.2016.08.005.
[3] Z. Shen, B. Qiao, L. Yang, W. Luo, X. Chen, Evaluating the influence of tooth surface wear on TVMS of planetary gear set, Mech. Mach. Theory 136
(2019) 206–223, doi:10.1016/j.mechmachtheory.2019.03.014.
[4] F. Chaari, W. Baccar, M.S. Abbes, M. Haddar, Effect of spalling or tooth breakage on gearmesh stiffness and dynamic response of a one-stage spur gear
transmission, Eur. J. Mech. - A/Solids 27 (2008) 691–705, doi:10.1016/j.euromechsol.20 07.11.0 05.
[5] X. Liang, M.J. Zuo, Z. Feng, Dynamic modeling of gearbox faults: a review, Mech. Syst. Signal Process. 98 (2018) 852–876, doi:10.1016/j.ymssp.2017.05.
024.
[6] J.H. Kang, Y.T. Yang, Estimate of mesh stiffness and load sharing ratio of a spur gear pair, in: Scottsdale, AZ, 1992: pp. 1–9.
[7] Y. Cai, Simulation on the rotational vibration of helical gears in consideration of the tooth separation phenomenon (a new stiffness function of helical
involute tooth pair), J. Mech. Des. 117 (1995) 460–469, doi:10.1115/1.2826701.
[8] D.C.H. Yang, J.Y. Lin, Hertzian Damping, Tooth friction and bending elasticity in gear impact dynamics, J. Mech. Trans. 109 (1987) 189, doi:10.1115/1.
3267437.
[9] X. Tian, Dynamic Simulation for System of Gearbox Including Localized Faults Master Thesis, University of Alberta, 2004.
[10] Z. Chen, Y. Shao, Dynamic simulation of spur gear with tooth root crack propagating along tooth width and crack depth, Eng. Fail. Anal. 18 (2011)
2149–2164, doi:10.1016/j.engfailanal.2011.07.006.
[11] Z. Chen, Y. Shao, Mesh stiffness calculation of a spur gear pair with tooth profile modification and tooth root crack, Mech. Mach. Theory 62 (2013)
63–74, doi:10.1016/j.mechmachtheory.2012.10.012.
[12] A. Fernandez del Rincon, F. Viadero, M. Iglesias, P. García, A. de-Juan, R. Sancibrian, A model for the study of meshing stiffness in spur gear transmis-
sions, Mech. Mach. Theory 61 (2013) 30–58, doi:10.1016/j.mechmachtheory.2012.10.008.
[13] H. Ma, J. Zeng, R. Feng, X. Pang, B. Wen, An improved analytical method for mesh stiffness calculation of spur gears with tip relief, Mech. Mach.
Theory 98 (2016) 64–80, doi:10.1016/j.mechmachtheory.2015.11.017.
[14] Z. Wan, H. Cao, Y. Zi, W. He, Z. He, An improved time-varying mesh stiffness algorithm and dynamic modeling of gear-rotor system with tooth root
crack, Eng. Fail. Anal. 42 (2014) 157–177, doi:10.1016/j.engfailanal.2014.04.005.
[15] T. Lin, H. Ou, R. Li, A finite element method for 3D static and dynamic contact/impact analysis of gear drives, Comput. Methods Appl. Mech. Eng. 196
(2007) 1716–1728, doi:10.1016/j.cma.2006.09.014.
[16] Y. Pandya, A. Parey, Failure path based modified gear mesh stiffness for spur gear pair with tooth root crack, Eng. Fail. Anal. 27 (2013) 286–296,
doi:10.1016/j.engfailanal.2012.08.015.
[17] I. Howard, S. Jia, J. Wang, The dynamic modeling of a spur gear in mesh including friction and a crack, Mech. Syst. Signal Process. 15 (2001) 831–853,
doi:10.10 06/mssp.20 01.1414.
[18] V.K. Ambarisha, R.G. Parker, Nonlinear dynamics of planetary gears using analytical and finite element models, J. Sound Vibration 302 (2007) 577–595,
doi:10.1016/j.jsv.2006.11.028.
[19] A. Fernández, M. Iglesias, A. de-Juan, P. García, R. Sancibrián, F. Viadero, Gear transmission dynamic: Effects of tooth profile deviations and support
flexibility, Appl. Acoust. 77 (2014) 138–149, doi:10.1016/j.apacoust.2013.05.014.
[20] L. Chang, G. Liu, L. Wu, A robust model for determining the mesh stiffness of cylindrical gears, Mech. Mach. Theory 87 (2015) 93–114, doi:10.1016/j.
mechmachtheory.2014.11.019.
[21] N.L. Pedersen, M.F. Jørgensen, On gear tooth stiffness evaluation, Comput. Struct. 135 (2014) 109–117, doi:10.1016/j.compstruc.2014.01.023.
[22] Y. Pandya, A. Parey, Experimental investigation of spur gear tooth mesh stiffness in the presence of crack using photoelasticity technique, Eng. Fail.
Anal. 34 (2013) 488–500, doi:10.1016/j.engfailanal.2013.07.005.
[23] J. Park, J. moon Ha, H. Oh, B.D. Youn, S. Park, J.-H. Choi, Experimental approach for estimating mesh stiffness in faulty states of rotating gear, in: 2015:
p. 7.
W. Chen, Y. Lei and Y. Fu et al. / Mechanism and Machine Theory 155 (2021) 104055 19

[24] R.G. Munro1, D. Palmer, L. Morrish, An experimental method to measure gear tooth stiffness throughout and beyond the path of contact, Proceed. Inst.
Mech. Eng., Part C: J. Mech. Eng. Sci. 215 (2001) 793–803, doi:10.1243/0954406011524153.
[25] N.K. Raghuwanshi, A. Parey, Experimental measurement of spur gear mesh stiffness using digital image correlation technique, Measurement 111 (2017)
93–104, doi:10.1016/j.measurement.2017.07.034.
[26] M. Feng, H. Ma, Z. Li, Q. Wang, B. Wen, An improved analytical method for calculating time-varying mesh stiffness of helical gears, Meccanica 53
(2018) 1131–1145, doi:10.1007/s11012-017-0746-6.
[27] C. Xie, L. Hua, J. Lan, X. Han, X. Wan, X. Xiong, Improved analytical models for mesh stiffness and load sharing ratio of spur gears considering structure
coupling effect, Mech. Syst. Signal Process. 111 (2018) 331–347, doi:10.1016/j.ymssp.2018.03.037.
[28] X. Liang, M.J. Zuo, M. Pandey, Analytically evaluating the influence of crack on the mesh stiffness of a planetary gear set, Mech. Mach. Theory 76
(2014) 20–38, doi:10.1016/j.mechmachtheory.2014.02.001.
[29] P. Sainsot and, P. Velex, O. Duverger, Contribution of gear body to tooth deflections—a new bidimensional analytical formula, J. Mech. Des. 126 (2004)
748–752, doi:10.1115/1.1758252.
[30] M.Ş. Tunalioğlu, B. Tuç, Theoretical and experimental investigation of wear in internal gears, Wear 309 (2014) 208–215, doi:10.1016/j.wear.2013.11.016.
[31] J.F. Archard, Contact and rubbing of flat surfaces, J. Appl. Phys. 24 (1953) 981–988, doi:10.1063/1.1721448.
[32] A. Flodin, S. Andersson, Simulation of mild wear in spur gears, Wear 207 (1997) 16–23, doi:10.1016/S0043- 1648(96)07467- 4.
[33] A. Flodin, S. Anderson, Wear simulation of spur gears, Tribotest 5 (1999) 25–50.
[34] A. Flodin, S. Andersson, Simulation of mild wear in helical gears, Wear 241 (20 0 0) 123–128, doi:10.1016/S0043-1648(00)00384-7.
[35] A. Flodin, S. Andersson, A simplified model for wear prediction in helical gears, Wear 249 (2001) 285–292, doi:10.1016/S0043-1648(01)00556-7.
[36] P. Bajpai, A. Kahraman, N.E. Anderson, A surface wear prediction methodology for parallel-axis gear pairs, J. Tribol. 126 (2004) 597–605, doi:10.1115/
1.1691433.
[37] H. Ding, Dynamic Wear Models for Gear System, The Ohio State University, 2007 PhD Thesis.
[38] J.A. Brandão, P. Cerqueira, J.H.O. Seabra, M.J.D. Castro, Measurement of mean wear coefficient during gear tests under various operating conditions,
Tribol. Int. 102 (2016) 61–69, doi:10.1016/j.triboint.2016.05.008.
[39] S. Feng, Analysis and calculation of gear mesh stiffness with tooth wear, JME 51 (2015) 27, doi:10.3901/JME.2015.15.027.
[40] Z. Shen, B. Qiao, W. Luo, L. Yang, X. Chen, The influence of external spur gear surface wear on the mesh stiffness, in: 2018 Prognostics and System
Health Management Conference (PHM-Chongqing), Chongqing, IEEE, 2018, pp. 1232–1238, doi:10.1109/PHM-Chongqing.2018.00216.
[41] X. Sun, T. Wang, R. Zhang, F. Gu, A.D. Ball, A model for mesh stiffness evaluation of spur gear with tooth surface wear, in: INSALyon, Université de
Lyon, 2019, p. 11.
[42] H. Liu, H. Liu, C. Zhu, Y. Zhou, A review on micropitting studies of steel gears, Coatings 9 (2019) 42, doi:10.3390/coatings9010042.
[43] Y. Huangfu, K. Chen, H. Ma, X. Li, X. Yu, B. Zhao, B. Wen, Investigation on meshing and dynamic characteristics of spur gears with tip relief under
wear fault, Sci. China Technol. Sci. 62 (2019) 1948–1960, doi:10.1007/s11431- 019- 9506- 5.
[44] V. Janakiraman, An Investigation of the Impact of Contact Parameters on the Wear Coefficient, The Ohio State University, 2013 Master Thesis.
[45] , Front matter, in: Beam Structures, John Wiley & Sons, Ltd, Chichester, UK, 2011, pp. i–xx, doi:10.1002/9781119978565.fmatter.
[46] H. Ma, R. Song, X. Pang, B. Wen, Time-varying mesh stiffness calculation of cracked spur gears, Eng. Fail. Anal. 44 (2014) 179–194, doi:10.1016/j.
engfailanal.2014.05.018.
[47] D.C.H. Yang, Z.S. Sun, A rotary model for spur gear dynamics, J. Mech., Transm. Autom. Des. 107 (1985) 529–535.
[48] S. Wen, P. Huang, Principles of Tribology, 2nd edition, Wiley, Hoboken, NJ, 2017.
[49] A. Flodin, Wear investigation of spur gear teeth, Tribotest 7 (20 0 0) 45–60, doi:10.10 02/tt.3020 070106.

You might also like