You are on page 1of 25

Mechanism and Machine Theory 175 (2022) 104958

Contents lists available at ScienceDirect

Mechanism and Machine Theory


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

Research paper

On the extended tooth contact and nonlinear dynamics for spur


gears—An analytical model
Xingyuan Zheng a , Yumei Hu a ,∗, Zao He a , Yue Xiao b , Xiangning Zhang c
a State Key Laboratory of Mechanical Transmission, Chongqing University, Chongqing 400044, China
b Applied Mechanics Laboratory, Department of Engineering Mechanics, Tsinghua University, 100084, Beijing, China
c
School of Mechanical Engineering and Automation, Chongqing Industry Polytechnic College, Chongqing, 401120, China

ARTICLE INFO ABSTRACT

Keywords: In gear mesh, the extended tooth contact is a natural reaction stemming from the elasticity of
Nonlinear dynamics solid bodies, which inevitably forces the theoretically separated teeth to engage and increases
Extended tooth contact the actual contact ratio. Consequently, the changed mesh stiffness waveforms will further
Analytical approach
influence the dynamic response. A lot of works are devoted to mesh stiffness representation
Mesh stiffness
and dynamic modeling. However, the correlation between extended tooth contact and nonlinear
Spur gears
dynamics still needs in-depth study. Motivated by a deeper understanding of the mesh
mechanisms and dynamics, this paper proposed an analytical model attempting to combine
the extended tooth contact and gear nonlinear dynamics. Starting from the gear elasticity,
an analytical mesh stiffness model is proposed, which extended the application ranges of the
popular fillet foundation polynomial formulas and covered both thin-walled and thick-walled
gears. Moreover, A load-sharing model is developed, which explicitly represented the correlation
between torques and intervals of extended tooth contact. Subsequently, the gear torsional
dynamic model is extended to embody the extended tooth contact. A series of comparisons
with FEM, existing theoretical models, and experiments demonstrated the acceptable accuracy
of the present model.

1. Introduction

Spur gears are key elements in mechanical transmission, widely applied in industrial, automotive, and aerial engineering. The
dynamic response of gears attracts a paramount concern due to the generation of noise, vibration, and harshness [1,2]. During
the oscillating, the internal excitation mainly comes from the mesh stiffness waveform, which induces complex responses as its
sensitivity in nonlinear dynamic equations [3–6]. Reliable predictions of gear dynamical behaviors greatly depend on the proper
representations of mesh stiffness.
In recent decades, the mesh stiffness modeling has become a central topic for geared systems. Prior methodologies including
experimental methods, finite element (FE) methods, and analytical models are widely applied in mesh stiffness evaluation.
Experimental methods [7–9] can directly achieve mesh stiffness but require dedicated and expensive devices in measurements.
Generally, FE methods are also reliable and accurate in mesh stiffness calculation [10]. Different discretization strategies are applied,
such as plane elements [11–13], solid elements [14–16], and finite element/contact mechanics model [17,18]. However, efficiencies
of FE methods are still limited by repeated pre-processing and large-scale computation.

∗ Corresponding author.
E-mail addresses: zhengxingyuan@cqu.edu.cn (X. Zheng), cdrhym@163.com (Y. Hu).

https://doi.org/10.1016/j.mechmachtheory.2022.104958
Received 3 March 2022; Received in revised form 30 April 2022; Accepted 23 May 2022
Available online 7 June 2022
0094-114X/© 2022 Elsevier Ltd. All rights reserved.
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Nomenclature

Alphabets

𝑏m Backlash
𝑐 Viscous damping
𝑒, 𝜖 Unloaded transmission errors
𝑓𝑚 , 𝑓𝑛 Mesh frequency, natural frequency
ℎ Ratio 𝑅𝑓 ∕𝑅int
𝑖 Imaginary unit: 𝑖2 = 1
𝑘 Indicator of loading conditions
𝑚 Tooth module
𝑚𝑒 Equivalent mass
𝑛 𝑛th Fourier transform
𝑞m Dynamic transmission error
(𝑟, 𝜃) Polar coordinates
𝑡 Time
𝑢𝐹 , 𝑢𝐹 1 , 𝑢𝐹 2 Moment arms
𝑢𝑟 (𝑅𝑓 , 𝜃) Radial displacement at outer boundary
𝑢𝜃 (𝑅𝑓 , 𝜃) Tangential displacement at outer boundary
𝑧 The number of tooth
𝐴rms rms of oscillating DTE component
A𝑛 , B𝑛 Fourier coefficients of stress, displacement components
𝐶𝑛 , 𝐷𝑛 Non-dimensional terms
𝐸 Young’s modulus
𝐹 Generic local mesh force
𝐹1 , 𝐹2 Local mesh force in tooth 1, 2
𝐽p , 𝐽g Rotary inertia of pinion, gear
𝐾𝑡 Tooth stiffness
𝐾ℎ Contact stiffness
𝐾𝑓 , 𝐾𝑓 12 , 𝐾𝑓 21 Fillet foundation stiffness
𝐾m Mesh stiffness
𝐾0 , 𝐾𝑛 Fourier contents of mesh stiffness
𝐿 Tooth width
𝑅𝑏 , 𝑅𝑓 Radius of base circle, root circle
𝑅int Inner bore radius
𝑆𝑓 Arc-length of junction
𝑇 Torque

Greeks

𝛼 Angular mesh position


𝛼𝐹 , 𝛼𝐹 1 , 𝛼𝐹 2 Action angles of 𝐹 , 𝐹1 , 𝐹2
𝛽1,2,3,4 Theoretical bounds of STC and DTC
𝛿, 𝛿1 , 𝛿2 Total elastic tooth pair deflections
𝛿𝑓 Foundation induced local tooth deflection
𝛿ℎ Contact compliance
𝜀∗ , 𝜀 Involute, extended contact ratio
𝜁 Damping ratio
𝜃𝑏 , 𝜃𝑓 Half tooth angles at base, root circles
𝜃p , 𝜃g Angular positions of pinion, gear
𝜅 Kolosov’s constant

Compared with experimental methods and FE methods, analytical methods are more flexible to reveal deeper insights on mesh
stiffness and thus attracts extensive attention [19]. Within the framework of analytical methods, the superposition principle is
commonly used. The flexibility of gear pair thus comprises the deformation of tooth, the contact compliances of tooth pairs, and the

2
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

𝜈 Poisson’s ratio
𝜌 Load sharing ratio
𝜎𝑟𝑟 (𝑅𝑓 , 𝜃) Radial stress at outer boundary
𝜎𝑟𝜃 (𝑅𝑓 , 𝜃) Tangential stress at outer boundary
𝜑1,2 Bounds of ETC
𝜙(𝛼) Angular tooth separation function
𝛹𝑠 , 𝛩𝑠1,2 , 𝛩𝑠1,2 Coefficients in asymptotic forms of 𝐾𝑓 , 𝐾𝑓 12 , 𝐾𝑓 21

Symbols

ℜ(∙) Real part


ℑ(∙) Imaginary part
N Piecewise function
inv Involute function
R Intervals
∥∙∥ Length of intervals

Abbreviations

DTC Double tooth contact


ETC Extended tooth contact
STC Single tooth contact
DTE Dynamic transmission error
FEM Finite element method
ICR Involute contact ratio
ECR Extended contact ratio
RK45 4–5th order Runge–Kutta method

elasticity of fillet foundation [20]. The tooth is typically regarded as a non-uniform cantilever beam. Accounting for the Hertzian
contact energy, bending energy and axial compressive energy, Yang et al. [21] proposed a potential energy method to calculate the
tooth deformation. Later, Tian et al. [22] modified the potential energy method by introducing the tooth shearing effect. In the model
of Yang [21], the interpenetration between tooth interfaces is linearized, which leads to the load-independency of contact stiffness.
To capture the nonlinearity of Hertzian contact, some scholars also adopted the load-dependent tooth contact, for example semi-
empirical formula [23] and bi-dimensional solutions [20]. Apart from the tooth deformation and contact compliance, the elasticity of
fillet foundation is also influences the mesh stiffness. Based on Muskhelishvili’s theory [24], Sainsot [25] derived the fillet foundation
induced local deflection and curve-fitted the solutions with polynomial formulas in plane strain. More recently, considering the
coupling flexibility of the adjacent teeth, Xie [26] further extended the polynomial formulas in plane strain, which largely improved
the accuracy in double tooth contact. Building upon the analytical framework, the mesh stiffness can be conveniently extended to
cover tooth failures [27–30], profile modifications [31–33], and manufacturing imperfections [34–36]. By inputting mesh stiffness
into dynamic equations in pre-specified forms [37–39] or instantaneously solved in each time step [40–42], the corresponding
dynamic behaviors can be estimated.
In these analytical models, intervals of mesh stiffness curve are commonly estimated with the theoretical tooth contact regions,
namely the involute contact. However, in gear mesh, a natural phenomenon named extended tooth contact (ETC) would occur. Due
to the flexibility of solid bodies, the applied torque would force theoretically separated teeth to engage. Consequently, the actual
contact regions are extended, which change the mesh stiffness waveforms and influence the subsequent dynamic responses [43].
Due to the complicated mesh mechanisms (e.g. coupling flexibility of adjacent teeth, tooth separations), to capture extended
tooth contact precisely and efficiently is a challenging job. On this complicated phenomenon, related studies typically based on
the polynomial formulas [25,26] or mixed with FE method. For example, Ma et al. [31] investigated the effects of extended tooth
contact on mesh stiffness and tip relief, where the fillet foundation stiffness is based on FE method and the Sainsot’s polynomial
formulas [25]. Xie et al. [44] adopted the extended polynomial formulas [26] and iteratively computed the extended contact region.
Admittedly, polynomial formulas [25,26] are in closed-form and can be conveniently integrated into coding. However, restrictions
still exist in these curve-fitted formulas, such as the confidence intervals and the case of plane stress for thin-walled gears [45].
On the other hand, prior studies mainly focus on the effects of extended tooth contact in quasi-statics, the quantitative correlation
between extended tooth contact and nonlinear dynamics still needs study in-depth.
Motivated by a deeper understanding of the mesh mechanisms and dynamics, this paper proposed an analytical model attempting
to bridge the extended tooth contact and gear nonlinear dynamics. Unlike the traditional analytical models, an extended fillet
foundation model is proposed, which broadened the application ranges the popular Sainsot’s and Xie’s polynomial formulas and
covered both thin-walled and thick-walled gears. A load-sharing model is further developed, which explicitly represented the

3
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

intervals of extended tooth contact and removed numerical iterations on the judgment of the extended contact region. Subsequently,
the gear torsional dynamic model is extended to embody the extended tooth contact. A series of comparisons with FEM, existing
models, and experiments demonstrated the acceptable accuracy of the present model. By controlling variables, effects of extended
tooth contact on both statics and dynamics are also investigated.
This paper is organized as follows: Section 2 introduced the kinematic model including stiffness components and load-
sharing; Section 3 detailed the modeling of nonlinear dynamics. Section 4 further realized the proposed models with a numerical
implementation. Solutions of the present model are further compared in Section 5. Parametric studies on extended tooth contact
are performed in Section 6. Algebras and parameters are summarized in Appendices.

2. Kinematics

2.1. Stiffness components

The flexibility of gear pair mainly stems from the beam actions of teeth, the contact compliances of tooth pairs, and the elasticity
of fillet foundation. For facilitating the subsequent discussions on the extended tooth contact and nonlinear dynamics, stiffness
components are separately modeled in the following subsections.

2.1.1. Tooth stiffness


The tooth is typically regarded as a non-uniform cantilever beam. The reaction of tooth to a generic local mesh force along the
line of action (LOA), denote as 𝐹 , is superposed by the beam actions under axial compression, bending, and shearing. Based on the
potential energy method [21,22], the tooth stiffness, denote as 𝐾𝑡 , write:
1 1 1 1
= + + , (2.1)
𝐾𝑡 𝐾𝑎 𝐾𝑏 𝐾𝑠
where 𝐾𝑎 , 𝐾𝑏 , and 𝐾𝑠 are axial compressive stiffness, bending stiffness, and shearing stiffness, respectively.
Parametric equations of 𝐾𝑎 , 𝐾𝑏 , and 𝐾𝑠 are given by:

1
𝜃𝑏 (𝜃𝑏 − 𝜏) cos 𝜏 sin2 𝛼𝐹
= 𝑑𝜏, (2.2)
𝐾𝑎 ∫−𝛼𝐹 2𝐸𝐿[sin 𝜏 + (𝜃𝑏 − 𝜏) cos 𝜏]
𝜃𝑏 3{1 + cos 𝛼 [(𝜃 − 𝜏) sin 𝜏 − cos 𝜏]}2 (𝜃 − 𝜏) cos 𝜏
1 𝐹 𝑏 𝑏
= 𝑑𝜏, (2.3)
𝐾𝑏 ∫−𝛼𝐹 2𝐸𝐿[sin 𝜏 + (𝜃𝑏 − 𝜏) cos 𝜏]3
𝜃𝑏 1.2(1 + 𝜈)(𝜃 − 𝜏) cos 𝜏 cos2 𝛼
1 𝑏 𝐹
= 𝑑𝜏, (2.4)
𝐾𝑠 ∫−𝛼𝐹 2𝐸𝐿[sin 𝜏 + (𝜃𝑏 − 𝜏) cos 𝜏]
where 𝛼𝐹 denotes the action angle of the local mesh force 𝐹 ; 𝜃𝑏 represents the half tooth angle at the base circle; 𝜏 ∈ [−𝛼𝐹 , 𝜃𝑏 ] is
an integral variable. 𝐸 denotes the Young’s modulus, and 𝐿 is the tooth width.

2.1.2. Tooth pair contact stiffness


The contact compliance of tooth pair would occur during the mesh. Many studies including FEM [13,18,46] and experiments [7]
indicate the load-dependency and nonlinearity of tooth contact. To capture this phenomenon, the semi-empirical formula [23] is
adopted:
𝐹 𝐸 0.9 𝐿0.8 𝐹 0.1
𝐾ℎ = = , (2.5)
𝛿ℎ 1.275
where 𝐾ℎ is the Hertzian contact stiffness of a generic tooth pair, and 𝛿ℎ is the corresponding contact compliance.

2.1.3. Fillet foundation stiffness


Apart from the deformation response of tooth, the flexibility of fillet foundation also contributes to the tooth deflection. Regarding
to the fillet foundation stiffness, polynomial formulas [25,26] are widely adopted. Admittedly, these formulas are in closed-form
and can be conveniently integrated into coding. However, restrictions still exist in these curve-fitted formulas:

• The confidence intervals of curve-fitted results are restricted in ℎ ∈ [1.4, 7], 𝜃𝑓 ∈ [0.01, 0.12] for Ref. [25] and ℎ ∈ [2, 5), 𝜃𝑓 ∈
[0.04, 0.08] for Ref. [26]. Here, ℎ is the ratio between root circle radius 𝑅𝑓 and inner bore radius 𝑅int , and 𝜃𝑓 is the half tooth
angle at root circle.
• Curve-fitted formulas [25,26] are restricted in the plane strain, while the plane stress is excluded. The practical engineering
employes both thick-walled and thin-walled gears extensively. According to Cornell [45], if the relation between the tooth
module 𝑚 and tooth width 𝐿 satisfies 2𝐿∕𝜋𝑚 > 5, the plane strain (e.g. thick-walled gears) is recommended; Otherwise, the
plane stress (e.g. thin-walled gears) is suggested.

4
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Fig. 2.1. Stress distribution at junction arc (outer boundary) in different loading conditions: (a) stress components caused by a generic local mesh force 𝐹
(𝑘 = 0); (b) stress components caused by the mesh force 𝐹1 (𝑘 = −1); (c) stress components caused by the mesh force 𝐹2 (𝑘 = +1).

Motivated by a wider scope of modern gear analysis and design, the fillet foundation stiffness is further modified in this section.
In the present model, both thick-walled gears and thin-walled gears are covered, while restrictions of ℎ and 𝜃𝑓 are removed. The
solution procedure will be detailed in Section 4.
Let us consider the fillet foundation (elastic ring) located in the center of polar coordinates (𝑟, 𝜃), where 𝑟 is the polar radius and
𝜃 is the polar angle. Fig. 2.1 covered three typical loading conditions for meshing gears, where 𝐹1 , 𝐹2 are mesh forces in tooth 1 and
2. 𝑢𝐹 , 𝑢𝐹 1 , and 𝑢𝐹 2 denote external moment arms of forces 𝐹 , 𝐹1 , and 𝐹2 . 𝑆𝑓 = 2𝜃𝑓 𝑅𝑓 represents the junction arc-length between
tooth and foundation.
In single tooth contact (see Fig. 2.1(a)), the displacement of foundation is only caused by the local force 𝐹 , which induces
radial stress components 𝜎𝑟𝑟 (𝑅𝑓 , 𝜃) and tangential stress components 𝜎𝑟𝜃 (𝑅𝑓 , 𝜃) at the outer boundary 𝑟 = 𝑅𝑓 . Situations are slightly
different in double tooth contact. As shown in Fig. 2.1(b)(c), due to the simultaneous tooth contact, additional stress components
will arise at the junction arc of adjacent teeth. Consequently, apart from the local displacement, additional displacements will occur.
An integer 𝑘 = 0, −1, +1 is further introduced to embody the aforementioned loading conditions in Fig. 2.1(a), (b), and (c)
respectively. As a result, the radial and tangential stress at outer boundary can be conveniently summarized as: 𝜎𝑟𝑟 (𝑅𝑓 , 𝜃 + 2𝜋𝑘∕𝑧),
𝜎𝑟𝜃 (𝑅𝑓 , 𝜃 + 2𝜋𝑘∕𝑧), where 𝑧 is the number of tooth. Inspired by [26], the cubic form of 𝜎𝑟𝑟 (𝑅𝑓 , 𝜃 + 2𝜋𝑘∕𝑧) and parabolic form of
𝜎𝑟𝜃 (𝑅𝑓 , 𝜃 + 2𝜋𝑘∕𝑧) are adopted, indicatively:
( ) ( )
⎧ 2𝜋𝑘 2𝜋𝑘 3
⎪𝜎 𝑟𝑟 𝑅 𝑓 , 𝜃 + = 𝜂1 𝜃 + + 𝜂2
𝑧 𝑧 (2.6)
⎨ ( ) ( )2 .
⎪𝜎 2𝜋𝑘 2𝜋𝑘
⎩ 𝑟𝜃 𝑅 𝑓 , 𝜃 + = 𝜂 3 𝜃 + + 𝜂4
𝑧 𝑧
Coefficients 𝜂𝑠 , 𝑠 = 1, 2, 3, 4 can be further determined according to the force and moment equilibrium condition:
𝜃𝑓
𝐹
𝐿 𝝈 ⋅ 𝐕d𝜃 = Φ, (2.7)
∫−𝜃𝑓 𝑅𝑓

the vector 𝝈 collects the stress components: 𝝈 = {𝜎𝑟𝑟 (𝑅𝑓 , 𝜃), 𝜎𝑟𝑟 (𝑅𝑓 , 𝜃)}. 𝐕 and Φ are dimensionless matrices, expressed as:
[ ] { }T
0 cos 𝜃 − sin 𝜃 𝑢 𝑢
𝐕= , Φ = cos 𝛼𝐹 (1 + 𝐹 ), − sin 𝛼𝐹 , − cos 𝛼𝐹 𝐹 . (2.8)
1 − sin 𝜃 − cos 𝜃 𝑅𝑓 𝑅𝑓

Based on the theory of Muskhelishvili [24], complex representations of stresses and displacements of elastic ring hold:
⎧ ( 2𝜋𝑘
) (
2𝜋𝑘
) ∑∞
⎪𝜎𝑟𝑟 𝑅𝑓 , 𝜃 + − 𝑖𝜎𝑟𝜃 𝑅𝑓 , 𝜃 + = A𝑛 exp(𝑖𝑛𝜃)
⎪ 𝑧 𝑧 𝑛=−∞
⎨ , (2.9)
⎪ 𝑅𝑓 (1 + 𝜈) ∑∞

⎪𝑢𝑟 (𝑅𝑓 , 𝜃) + 𝑖𝑢𝜃 (𝑅𝑓 , 𝜃) = B𝑛 exp(𝑖𝑛𝜃)


⎩ 𝐸 𝑛=−∞

where 𝑢𝑟 (𝑅𝑓 , 𝜃) and 𝑢𝜃 (𝑅𝑓 , 𝜃) are the corresponding radial and tangential displacements at the outer boundary 𝑟 = 𝑅𝑓 . 𝑖 denotes the
unit imaginary number, namely 𝑖2 = −1, and 𝜈 represents the Poisson’s ratio. A𝑛 and B𝑛 are the 𝑛 − th harmonic coefficients of the

5
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Fourier series for stress and displacement components, expressed as:


𝜋[ ( ) ( )]
⎧A = 1 𝜎 𝑅𝑓 , 𝜃 +
2𝜋𝑘
− 𝑖𝜎𝑟𝜃 𝑅𝑓 , 𝜃 +
2𝜋𝑘
exp(−𝑖𝑛𝜃)d𝜃
⎪ 𝑛 2𝜋 ∫−𝜋 𝑟𝑟 𝑧 𝑧
⎨ 𝜋[ ] . (2.10)
⎪B = 1
⎩ 𝑛 2𝜋 ∫−𝜋 𝑢𝑟 (𝑅𝑓 , 𝜃) + 𝑖𝑢𝜃 (𝑅𝑓 , 𝜃) exp(−𝑖𝑛𝜃)d𝜃

If displacements at the inner bore 𝑟 = 𝑅int are fixed, the linear relation between A𝑛 and B𝑛 writes [25,26]:
B𝑛 = 𝐶𝑛 (𝜅, ℎ)A𝑛 + 𝐷𝑛 (𝜅, ℎ)A−𝑛 , (2.11)

where 𝐶𝑛 (𝜅, ℎ) and 𝐷𝑛 (𝜅, ℎ) are non-dimensional terms detailed in Appendix A. 𝜅 is the Kolosov’s constant, defined as 𝜅 = 3 − 4𝜈 for
plane strain and 𝜅 = (3 − 𝜈)∕(1 + 𝜈) for plane stress. A−𝑛 is the conjugate of A−𝑛 .
Combining Eqs. (2.9)–(2.10) and letting 𝜃 = 0, yields the triplet displacement components of fillet foundation at 𝑟 = 𝑅𝑓 , 𝜃 = 0:
{ ∞ }
𝑅𝑓 (1 + 𝜈) ∑ ∑∞ ∑∞
𝐮= ℜ(B𝑛 ), ℑ(B𝑛 ), 𝑛ℜ(𝑖B𝑛 ) . (2.12)
𝐸 𝑛=−∞ 𝑛=−∞ 𝑛=−∞
{ }
In Eq. (2.12), the vector 𝐮 = 𝑢𝑟 (𝑅𝑓 , 0), 𝑢𝜃 (𝑅𝑓 , 0), 𝜕𝑢𝑟 (𝑅𝑓 , 0)∕𝜕𝜃 . The symbol ℜ(∙) represents the real part of ∙, while ℑ(∙) denotes
the imaginary part of ∙.
According to Sainsot [25], the local tooth deflection contributed by the foundation deformation, denotes as 𝛿𝑓 , can be treated
as the corresponding transport motion of rigid tooth along LOA, expressed as:

𝛿𝑓 = 𝐮 ⋅ ΦT . (2.13)

Substituting 𝑘 = 0, −1, +1 into A𝑛 in Eq. (2.9) and combining Eqs. (2.10)–(2.13), gives asymptotic forms of fillet foundation
stiffness components:
2 [( ) ]
⎧ 1 cos2 𝛼𝐹 ∑ 𝑢𝐹 𝑠
⎪ = 𝛹𝑠
⎪ 𝐾𝑓 𝐸𝐿 𝑠=0 𝑆𝑓
⎪ ( )
⎪ 1 cos 𝛼𝐹 1 cos 𝛼𝐹 2 2,1 2,1 𝑢𝐹 1 2,1 𝑢𝐹 2 2,1 𝑢𝐹 1 𝑢𝐹 2
⎨𝐾 = 𝛩0 + 𝛩1 + 𝛩2 + 𝛩3 , (2.14)
⎪ 𝑓 21 𝐸𝐿 𝑆𝑓 𝑆𝑓 𝑆𝑓2
⎪ ( )
⎪ 1 cos 𝛼𝐹 1 cos 𝛼𝐹 2 𝑢 𝑢 𝑢 𝑢
⎪𝐾 = 𝛩01,2 + 𝛩11,2 𝐹 2 + 𝛩21,2 𝐹 1 + 𝛩31,2 𝐹 1 𝐹 2
⎩ 𝑓 12 𝐸𝐿 𝑆 𝑓 𝑆 𝑓 𝑆𝑓2
where 𝐾𝑓 , 𝐾𝑓 21 , and 𝐾𝑓 12 are components of fillet foundation stiffness. 𝐾𝑓 refers to the deformation resistance to the local mesh
force 𝐹 (see Fig. 2.1(a)); 𝐾𝑓 21 refers to the deformation resistance of tooth 2 to the mesh force 𝐹1 (see Fig. 2.1(b)); Similarly, 𝐾𝑓 12
refers to the deformation resistance of tooth 1 to the mesh force 𝐹2 (see Fig. 2.1(c)). It can be observed that terms 𝐾𝑓 , 𝐾𝑓 21 , and
𝐾𝑓 12 are controlled by the load independent infinite series 𝛷𝑠 , 𝛩𝑠1,2 and 𝛩𝑠2,1 , which are detailed in Appendix B for concise.

2.2. Load-sharing model

2.2.1. Load–deflection constitutive relation


For a pair of mating gears, we name the driving part (applied with torque 𝑇 ) as the pinion, and driven part is the gear. During
the mesh, deflections of tooth pairs change with the varying engagement position. For convenience, the force angle of pinion tooth
1 (namely 𝛼𝐹 1,p , abbreviated as 𝛼𝐹 1,p = 𝛼) is selected as the reference engagement position.
Recalling the stiffness components in Eqs. (2.1), (2.5), and (2.14), gives the load–deflection constitutive relation:
( )
⎧ ∑ 1 1 1 ∑ 1
⎪𝛿1 (𝛼, 𝐹1 , 𝐹2 ) = 𝐹1 + + + 𝐹2
⎪ 𝑗=p,g
𝐾𝑡1,𝑗 𝐾𝑓 11,𝑗 2𝐾ℎ1 𝐾
𝑗=p,g 𝑓 12,𝑗
⎨ ( ) , (2.15)
∑ 1 1 1 ∑ 1
⎪𝛿 (𝛼, 𝐹 , 𝐹 ) = 𝐹 + + + 𝐹
⎪ 2 1 2 2
𝐾𝑡2,𝑗 𝐾𝑓 22,𝑗 2𝐾ℎ2 1
𝐾
⎩ 𝑗=p,g 𝑗=p,g 𝑓 21,𝑗

In Eq. (2.15), 𝛿(𝛼, 𝐹1 , 𝐹2 ) denotes the total elastic deflection of a certain tooth pair. Subscripts ∙1 and ∙2 refer to the previously
defined quantities in tooth pair 1 and 2. Similarly, subscripts ∙p and ∙g refer to the quantities of pinion and gear, respectively. 𝑗 is
an indicator with 𝑗 = p or 𝑗 = g. 𝐾𝑓 11 and 𝐾𝑓 22 are the local fillet foundation stiffness of tooth 1 and tooth 2, respectively.

2.2.2. Intervals of tooth contact modes


This section prescribed intervals of tooth contact modes for facilitating the later discussions. Fig. 2.2 illustrated the gear pair
arrangement. In the initial position (see Fig. 2.2(a)), the tooth pair 1 is aligned at the pitch point, while the tooth pair 2 has yet to
engage. As shown in see Fig. 2.2(b), the situation will be reversed after one mesh cycle, where the tooth pair 2 will engage at the
pitch point, and tooth pair 1 has already separated.
Within this specified mesh cycle, theoretical intervals for single tooth contact (STC) and double tooth contact (DTC) can be readily
defined as: 𝛼 ∈ [𝛽1 , 𝛽2 ) for the first STC, 𝛼 ∈ [𝛽2 , 𝛽3 ] for the DTC, and 𝛼 ∈ (𝛽3 , 𝛽4 ] for the second STC (see dashed lines in Fig. 2.2(c)).

6
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Fig. 2.2. Gear pair arrangement: (a) the initial position of mounted gear pair; (b) the final position of mounted gear pair; (c) expected mesh stiffness curves
with intervals of single tooth contact (STC), extended tooth contact (ETC), and double tooth contact (DTC). (For interpretation of the references to color in this
figure legend, the reader is referred to the web version of this article.)

However, in practice, the deformation of tooth pairs will extend the contact duration, which speeds up the tooth engaging-in and
delays the tooth engaging-out. This phenomenon is often called the extended tooth contact (ETC) [31]. If the extended tooth contact
(ETC) is considered, smooth transitions between STC and DTC will occur (see blue lines in Fig. 2.2(c)).
Base on the above descriptions, intervals of tooth contact modes are prescribed as:

RSTC = [𝛽1 , 𝜑1 ) ∪ (𝜑2 , 𝛽4 ], RDTC = [𝛽2 , 𝛽3 ], RETC = [𝜑1 , 𝛽2 ) ∪ (𝛽3 , 𝜑2 ], (2.16)

where RSTC , RDTC , RETC denote regions of STC, DTC, and ETC, respectively. According to the involute geometry, expressions of 𝛽1 ,
𝛽2 , 𝛽3 , and 𝛽4 are given in Appendix C. 𝜑1 and 𝜑2 defined the interval of ETC, which will be detailed in Section 2.2.4.

2.2.3. Double tooth contact


Essentially, during simultaneous engagements, the deformation compatibility of tooth pairs determins the load sharing. If the
deformed tooth pairs are still in contact, the pitch length variation of pinion and gear must be equal. According to the equivalent
spring model by Ref. [47], the deformation compatibility in DTC holds:
∑ [ ]
(𝛿2,𝑗 − 𝛿1,𝑗 ) + (𝑒2,𝑗 − 𝑒1,𝑗 ) = 0, (2.17)
𝑗=p,g

where 𝛿1,𝑗 and 𝛿2,𝑗 represent deflection components of tooth 1 and 2 in pinion (𝑗 = p) and gear (𝑗 = g); 𝑒1,𝑗 and 𝑒2,𝑗 are the
corresponding unloaded transmission errors. For concise, we let 𝜖 = (𝑒2,𝑔 + 𝑒2,𝑝 ) − (𝑒1,𝑔 + 𝑒1,𝑝 ). Obviously, 𝛿𝑠,𝑗 and 𝛿𝑠 (𝐹1 , 𝐹2 ) satisfies:

𝛿𝑠 (𝛼, 𝐹1 , 𝐹2 ) = 𝛿𝑠,𝑗 , 𝑠 = 1, 2. (2.18)
𝑗=p,g

Combining Eqs. (2.15), (2.17), and (2.18), the load sharing in DTC can be determined by:
⎧𝛿 (𝛼, 𝐹 , 𝐹 ) − 𝛿 (𝛼, 𝐹 , 𝐹 ) + 𝜖 = 0
⎪ 2 1 2 1 1 2
⎨𝐹 + 𝐹 = 𝑇 , 𝛼 ∈ RDTC , (2.19)
⎪ 1 2
𝑅𝑏,p

where 𝑅𝑏,p is the base circle radius of pinion.

2.2.4. Extended tooth contact


Strictly speaking, the extended tooth contact (ETC) is a special case of double tooth contact (DTC). However, in extended tooth
contact (ETC), the theoretically separated tooth pair will be pressed to engage due to the flexibility of tooth pairs. Consequently,
the deformation compatibility in Eq. (2.19) should be extended to embody the tooth separation distance.
If small deflection holds, the tooth separation can be regarded as additional tooth errors of pinion teeth (the driving part). By
denoting 𝜙(𝛼) as the angular tooth separation distance, the deformation compatibility in the extended tooth contact (ETC) becomes:

[ ]
⎧ 1 𝛿2 (𝛼, 𝐹1 , 𝐹2 ) − 𝛿1 (𝛼, 𝐹1 , 𝐹2 ) + 𝜖 = 𝜙(𝛼)
⎪ 𝑅𝑏,p
⎨ , 𝛼 ∈ RETC , (2.20)
⎪𝐹 1 + 𝐹 2 = 𝑇
⎩ 𝑅𝑏,p
Detailed expression of the angular tooth separation distance function 𝜙 see Ref. [48] (see also Appendix D for convenience). From
now, the remaining target is to solve the boundary of RETC , namely 𝜑1 and 𝜑2 . As shown in Fig. 2.3(a)(b), 𝜑1 and 𝜑2 refer to two
limiting cases.

7
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Fig. 2.3. Limiting cases in ETC: (a) deformed tooth pair 1 in engaging-out condition, 𝐹1 = 0; (b) deformed tooth pair 2 in engaging-in condition, 𝐹2 = 0.

For the first limiting case in Fig. 2.3(a), the outgoing rigid tooth pair 1 has already separated, while the deformed tooth pair 1
is just at the end of engagement. As a result, when 𝛼 = 𝜑1 , the mesh force of the deformed tooth pair 1 is 𝐹1 = 0, and the deformed
system reached the onset of simultaneous tooth contact. Whereas, the rigid system is still in STC until 𝛼 = 𝛽2 . Therefore, the so-called
extended tooth contact reduced the theoretical STC region by 𝛽2 − 𝜑1 . Analogously, for the second limiting case in Fig. 2.3(b), the
incoming rigid tooth pair 2 has not engaged yet, while the deformed tooth pair 2 has just engaged. When 𝛼 = 𝜑2 , the mesh force
of deformed tooth pair 2 is 𝐹2 = 0, and the theoretical STC region is reduced by 𝜑2 − 𝛽3 .
Separately substituting 𝐹1 = 0 or 𝐹2 = 0 into Eq. (2.20) gives implicit functions for 𝜑1 and 𝜑2 :
⎧ 1 [ ( 𝑇
) (
𝑇
) ]
⎪ 𝛿2 𝜑1 , 0, − 𝛿1 𝜑1 , 0, + 𝜖 = 𝜙(𝜑1 )
⎪ 𝑅𝑏,p 𝑅𝑏,p 𝑅𝑏,p
⎨ [ ( ) ( ) ] , (2.21)
⎪ 1 𝛿 𝜑 , 𝑇 , 0 − 𝛿 𝜑 , 𝑇 , 0 + 𝜖 = 𝜙(𝜑 )
⎪ 𝑅𝑏,p 2 2
𝑅𝑏,p 1 2
𝑅𝑏,p 2

Eq. (2.21) implicitly embodied the correlation between the bounds of ETC and the torque 𝑇 , which can be numerically solved.
Interestingly, based on the small deflection assumption, the small change of engagement position has negligible effect on the
stiffness components, for example: 𝐾𝑓 (𝛼 + 𝛥𝛼) ≈ 𝐾𝑓 (𝛼), 𝐾𝑡 (𝛼 + 𝛥𝛼) ≈ 𝐾𝑡 (𝛼). With this consideration, explicit expressions for 𝜑1 and
𝜑2 can be obtained. Noticing 𝜙(𝛽2 ) = 𝜙(𝛽3 ) = 0 and combining Eqs. (2.15), (2.16), (2.21) yields:
[ ]
⎧ 𝑇 ∑ 1 1 1 1 𝜖
⎪𝛽2 − 𝜑1 ≈ 2 + + − +
⎪ 𝑅𝑏,𝑝 𝑗=𝑝,𝑔 𝐾𝑡2,𝑗 (𝛽2 ) 𝐾𝑓 22,𝑗 (𝛽2 ) 2𝐾ℎ2 (𝑇 ) 𝐾𝑓 12,𝑗 (𝛽2 ) 𝑅𝑏,𝑝
⎨ [ ] (2.22)
∑ 1 1 1 1
⎪𝜑 − 𝛽 ≈ 𝑇 + + − +
𝜖
⎪ 2 3 2
𝑅𝑏,𝑝 𝑗=𝑝,𝑔 𝐾𝑡1,𝑗 (𝛽3 ) 𝐾𝑓 11,𝑗 (𝛽3 ) 2𝐾ℎ1 (𝑇 ) 𝐾𝑓 21,𝑗 (𝛽3 ) 𝑅𝑏,𝑝

The extended tooth contact also attracts many recent researches [31,44]. Compared with these works, Eq. (2.22) explicitly expressed
the load-dependency of ETC intervals and removed iterations on distinguishing DTC and ETC, which facilitates the inclusion of this
phenomenon in the subsequent dynamic modeling and parametric analysis.

2.2.5. Involute contact ratio and extended contact ratio


By definition, the calculation of the involute contact ratio (ICR), denote as 𝜀∗ , follows:
2 ∥ RDTC ∥ + ∥ R∗STC ∥ ∥ RDTC ∥
𝜀∗ = =1+ , (2.23)
∥ R0 ∥ ∥ R0 ∥
where the parallel symbol ∥ ∙ ∥ represents the length of interval. ∥ R0 ∥= 𝛽4 − 𝛽1 denotes the duration of the mesh cycle defined in
Section 2.2.2; ∥ R∗STC ∥= (𝛽2 − 𝛽1 ) + (𝛽4 − 𝛽3 ) is the length of theoretical STC. ∥ RDTC ∥= 𝛽3 − 𝛽2 is the length of DTC.
The involute contact ratio (ICR) greatly influences the gear dynamics [49], which is a key parameter in the gear design. However,
as discussed in Sections 2.2.3–2.2.4, due to the extended tooth contact, the actual contact ratio will be increased. For a deeper insight
on this phenomenon, the contact ratio considering the extended tooth contact is analogously defined as:
( )
2 ∥ RDTC ∥ + ∥ RETC ∥ + ∥ RSTC ∥ ∥ RDTC ∥ + ∥ RETC ∥
𝜀(𝑇 ) = =1+ , (2.24)
∥ R0 ∥ ∥ R0 ∥
where 𝜀(𝑇 ) is a load-dependent coefficient, named as the extended contact ratio (ECR). ∥ RETC ∥= (𝛽2 − 𝜑1 ) + (𝜑2 − 𝛽3 ) is the length
of ETC, which can be evaluated according to Eq. (2.22).

8
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Fig. 3.1. 2DOF torsional gear dynamical model with gear compliance.

3. Nonlinear dynamics

3.1. Mesh stiffness waveforms

In sense of dynamics, the mesh stiffness waveforms are the main sources of internal excitation. In light of the load sharing model
proposed in Section 2.2, mesh forces 𝐹1 , 𝐹2 now become known parameters. For a certain mesh position 𝛼 ∈ R0 , the mesh stiffness
can be determined with the corresponding mesh forces and tooth deflections:
⎧ 𝐹
, 𝛼 ∈ RSTC
⎪ 𝛿(𝛼, 𝐹1 , 𝐹2 )
𝐾m0 (𝛼, 𝑇 ) = ⎨ , (3.1)
𝐹1 𝐹2
⎪ + , 𝛼 ∈ RETC ∪ RDTC
⎩ 𝛿1 (𝛼, 𝐹1 , 𝐹2 ) 𝛿2 (𝛼, 𝐹1 , 𝐹2 )
Eq. (3.1) evaluated the mesh stiffness in the specified mesh cycle, namely 𝛼 ∈ R0 . 𝐾m0 (𝛼, 𝑇 ) is a load-dependent term denoting the
mesh stiffness of gear pair when 𝛼 ∈ R0 . Stiffness curves in the subsequent mesh cycles (𝛼 ∈ R𝑠 , 𝑠 = 1, 2, 3, …) can be analogously
mirrored and assembled due to the periodic engagements of tooth pairs. Consequently, the time-varying waveform of mesh stiffness
curve becomes:
{ }
𝐾m (𝑡, 𝑇 ) = 𝐾m𝑠 (𝛼, 𝑇 ) 𝑠=0,1,2,3,… (3.2)

where 𝐾m (𝑡, 𝑇 ) is the time-varying waveform of mesh stiffness curve, and 𝑡 is the time. 𝐾m𝑠 (𝛼, 𝑇 ) denotes the mesh stiffness the mesh
duration R𝑠 . A Fourier representation of 𝐾m (𝑡, 𝑇 ) holds:


𝐾m (𝑡, 𝑇 ) = 𝐾0 (𝜀, 𝑇 ) + 𝐾𝑛 (𝜀, 𝑇 ) cos(2𝜋𝑛𝑓m 𝑡 − 𝜓𝑛 ). (3.3)
𝑛=1

Same to Eq. (2.9), the integer 𝑛 denotes the order of Fourier transforms. 𝐾0 is the average mesh stiffness value. 𝐾𝑛 and 𝜓𝑛 are the
𝑛th Fourier coefficient and phase angle, respectively. 𝑓m is the mesh frequency.
Eq. (3.3) emphasized the influences of torque 𝑇 on mesh stiffness. The nonlinear Hertzian contact and structural coupling of
fillet foundation lead to the load-dependency of harmonic contents 𝐾𝑛 (𝜀, 𝑇 ). On the other hand, as discussed in Section 2.2, the
torque 𝑇 further extends the tooth contact and increases the contact ratio. As a result, harmonic contents 𝐾𝑛 (𝜀, 𝑇 ) are also functions
of the extended contact ratio 𝜀.

3.2. Equations of motion

For the dynamic modeling, basic assumptions of torsional models with gear compliance [50] are adopted. A two-degrees-of-
freedom model of a pair of spur gear system is illustrated in Fig. 3.1. 𝐽p and 𝐽g are rotary inertias of pinion and gear, respectively.
𝑇p and 𝑇g are the corresponding torques. 𝑐 is the viscous damping in gear mesh. 𝜖(𝑡) is an equivalent displacement excitation, namely
unloaded transmission error.
If effects of shaft rigidity and friction are negligible, equations of motion for the spur gear pair hold:
[ ( )]
𝑑 2 𝜃p (𝑡) 𝑑𝑞m (𝑡) 𝑑𝜖(𝑡)
𝐽p = 𝑇p − 𝑅𝑏,p 𝐾m (𝑡, 𝑇 )N (𝑞m (𝑡)) + 𝑐 − ;
𝑑𝑡2 𝑑𝑡 𝑑𝑡
[ ( )] , (3.4)
𝑑 2 𝜃g (𝑡) 𝑑𝑞m (𝑡) 𝑑𝜖(𝑡)
𝐽g = 𝑅𝑏,g 𝐾m (𝑡, 𝑇 )N (𝑞m (𝑡)) + 𝑐 − − 𝑇g ,
𝑑𝑡2 𝑑𝑡 𝑑𝑡
where 𝜃𝑝 and 𝜃𝑔 represent the angular positions of pinion and gear, respectively; Again, 𝑅𝑏,p and 𝑅𝑏,g are base circle radius of pinion
and gear. 𝑞m (𝑡) is the dynamic transmission error (DTE), expressed as:

𝑞m (𝑡) = 𝜃p (𝑡)𝑅𝑏,p + 𝜃g (𝑡)𝑅𝑏,g . (3.5)

9
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Fig. 4.1. Flowchart of mesh stiffness evaluation and dynamic computation. (For interpretation of the references to color in this figure legend, the reader is
referred to the web version of this article.)

Letting 2𝑏m be the total backlash, yields the function N (𝑞m (𝑡)):
⎧𝑞 (𝑡) − 𝜖(𝑡) − 𝑏 𝑏𝑚 < 𝑞m (𝑡)
⎪ m m
N (𝑞m (𝑡)) = ⎨0 ∣ 𝑞m (𝑡) ∣≤ 𝑏m . (3.6)
⎪𝑞m (𝑡) − 𝜖(𝑡) + 𝑏𝑚 𝑞m (𝑡) < −𝑏m

Recalling Eq. (3.3) and neglecting torque fluctuations and slight tooth imperfections, equations of motion in Eq. (3.4) becomes:
[ ]
d2 𝑞m (𝑡) 𝑑𝑞m (𝑡) ∑𝑁
𝑇
𝑚𝑒 +𝑐 + 𝐾0 (𝜀, 𝑇 ) + 𝐾𝑛 (𝜀, 𝑇 ) cos(2𝜋𝑛𝑓𝑚 𝑡 − 𝜓𝑛 ) N (𝑞m (𝑡)) = , (3.7)
𝑑𝑡2 𝑑𝑡 𝑛=1
𝑅𝑏,p

where 𝑚𝑒 = 𝐽p 𝐽g ∕(𝐽g 𝑅2𝑏,p + 𝐽p 𝑅2𝑏,g ) is equivalent mass. In the present model, we let the integer 𝑁 = 4, which is sufficiently accurate
to capture the dynamic resonances. For the consistency in static and dynamic modeling, the steady torque is applied to the pinion
𝑇p = 𝑇 , and the steady velocity is applied to the gear. Similar to Ref. [51], the viscous damping 𝑐 is determined by the damping
ratio 𝜁 and the average mesh stiffness 𝐾0 (𝜀, 𝑇 ):

𝑐 = 2𝜁 𝑚𝑒 𝐾0 (𝜀, 𝑇 ) (3.8)
Due to the discontinuity of the elastic mesh force and the nonlinearity of mesh stiffness, the dynamic equation in Eq. (3.7) is
numerically solved. The systematic framework throughout this paper which will be detailed in Section 4.

4. Numerical implementation

4.1. Framework

In light of the kinematic modeling in Section 2 and the dynamic modeling in Section 3, flowchart of mesh stiffness evaluation
and dynamic computation can be modularized in Fig. 4.1. The coding platform is Matlab® .
The inputting procedure is user-friendly, which only needs the necessary gear pair parameters and driving torques 𝑇 . Based on
the Cornell criterion [45], the program adaptively selects the Kolosov’s constant 𝜅 to cover both thick-walled gears (plane strain)
and thin-walled gears (plane stress). To reduce the computational expense, the prescribed interval R0 is uniformly discretized:
{ }
(𝛽4 − 𝛽1 ) 2(𝛽4 − 𝛽1 ) 𝑠(𝛽4 − 𝛽1 )
𝛼 ∈ R0 = , ,…, ,… , (4.1)
𝑃 𝑃 𝑃 𝑠=1,2,3,…,𝑃

where the integer 𝑃 controls the resolution. For the present analysis, we let the integer 𝑃 = 100. Consequently, matrix operation
can be applied throughout the subsequent load-sharing and mesh stiffness evaluation.
The evaluation of mesh stiffness comprises the preparations of stiffness components and solutions of load-sharing. Benefiting
from the analytical forms of stiffness components (Eqs. (2.1), (2.5), and (2.14)) and explicit representations of ETC interval

10
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Fig. 4.2. Solutions of infinite series of pinion tooth 1 in gear pair B (ℎ = 10): (a) 𝛼 = 10◦ ; (b) 𝛼 = 20◦ .

(Eq. (2.22)), the calculation of mesh stiffness becomes straightforward and efficient. As a result, the intervals RSTC , RDTC , and
RETC are automatically determined and no additional numerical iterations are required (see red dashed box in Fig. 4.1).
It worth remarking that the calculation of infinite series of fillet foundation stiffness should be carefully considered (see blue
dashed box in Fig. 4.1). For concise, vectors Ψ𝑛 , Θ1,2 2,1
𝑛 , Θ𝑛 are used to collect the corresponding infinite series in each accumulation
step 𝑛, indicatively:
{ } { } { }
1,2 1,2 1,2 1,2 2,1 2,1 2,1 2,1
Ψ𝑛 = 𝛹0 , 𝛹1 , 𝛹2 , Θ1,2
𝑛 = 𝛩0 , 𝛩1 , 𝛩2 , 𝛩3 , Θ2,1
𝑛 = 𝛩0 , 𝛩1 , 𝛩2 , 𝛩3 . (4.2)

The differences of vectors Ψ𝑛 , Θ1,2 2,1


𝑛 , Θ𝑛 between each accumulation steps 𝑛 and 𝑛 − 1 are recorded. The maximum value of the
difference, denote as 𝜉, is:
{ }
1,2 2,1
𝜉 = max |Ψ𝑛 − Ψ𝑛−1 |, |Θ1,2 2,1
𝑛 − Θ𝑛−1 |, |Θ𝑛 − Θ𝑛−1 | . (4.3)

Once 𝜉 reaches 1e−6, infinite series can be considered as solved. During the calculation of infinite series, 𝐶±𝑛 (𝜅, ℎ) and 𝐷±𝑛 (𝜅, ℎ)
∗ (𝜅, ℎ) and 𝐷∗ (𝜅, ℎ) (see Eq. (A.3) in Appendix A), which
in Eq. (2.11) are recommend to input as fractional splitting forms 𝐶±𝑛 ±𝑛
efficiently avoids the numerical problem ‘‘inf ∗ 0 = nan’’.
To further illustrate the solution procedure of Ψ𝑛 , Θ1,2 2,1
𝑛 , Θ𝑛 , an example for pinion tooth 1 in gear pair B (properties see
Appendix E) is shown in Fig. 4.2. In this example, the mesh positions are 𝛼 = 10, 20◦ . The value of ℎ is selected to be ℎ = 10, which
already exceeds confidence intervals of Ref. [25] (ℎ ∈ [1.4, 7]) and Ref. [26] (ℎ ∈ [2, 5)). Beyond confidence intervals of Refs. [25,26],
a good convergence and arithmetic stability of the present method are still observed: curves of all infinite series slightly fluctuated
in the initial phases and gradually leveled off after short accumulation times.
In the present model, the evaluation of mesh stiffness is a priori in the dynamic analysis. Based on the calculated mesh stiffness
waveforms 𝐾m (𝑡, 𝑇 ), the dynamic equation Eq. (3.7) is further integrated through 4-5th order Runge–Kutta method (RK45). The
solution strategy is adopted from our previous work [46] (see also black dashed box in Fig. 4.1, pictured below). The stepping-
up and stepping-down processes are applied to capture the bifurcation of gear dynamic response. In the stepping-up process, the
temporary frequency 𝑓𝑠 starts at the initial mesh frequency 𝑓initial , which is gradually increased to the final mesh frequency 𝑓f inal .
In the stepping-down process, such procedure is reversed.

4.2. Computational expenses

The computation time is estimated with the function ‘tic-toc’ in Matlab® . All numerical computations are realized on a 4 GHz
quad-core CPU with 32 GB of RAM. Computational expenses are summarized as follows:

• In the module of the mesh stiffness calculation in one mesh cycle (100 data points), the average computation time is 1.33 s.
• In the module of solving nonlinear dynamic equations (600 stepping-up/down), the average cost is 4.48 s.

Compared with the FEM [13,46] in our previous works, the computational expenses on mesh stiffness are greatly reduced. On the
other hand, compared with polynomial formulas [25,26], the present model extended the application range to cover both thick-
walled gears (plane strain) and thin-walled gears (plane stress), while remains acceptable computational efficiency for the further
analysis and optimization designs.

11
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Fig. 5.1. Configurations of FE models for gear pair A.

5. Model verification

Verifications are performed in this section, including deformation responses (see Section 5.1) and dynamics (see Section 5.2).
Selections of comparison objects are highlighted as follows:

• In the comparisons of deformation responses, two standard spur gear pairs are selected, labeled as gear pair A (ICR = 1.71)
and gear pair B (ICR = 1.75). According to the Cornell criterion [45], gear pair A is thick-walled gear satisfying the plane
strain, while gear pair B is thin-walled gear satisfying the plane stress.
• For comparisons of dynamics, the well-known experimental results of A. Kahraman et al. [49,52] are adopted. The gear pair
B with involute contact ratio ICR = 1.37, 1.75, and 1.77 are included.

For clarity, geometrical and material properties of gear pair A and B are summarized in Appendix E.

5.1. Deformation responses

Solutions of load-sharing and mesh stiffness are compared in this section. FE models are specially designed for the verification.
The strategy of geometry discretization is adopted from our previous works [13,46]. The simulation platform is Abaqus® . 2D
elements for gear pair A is CPE4R (plane strain) and for gear pair B is CPS4R (plane stress). An example for the FE model of
gear pair A is shown in Fig. 5.1.
In FE modeling, contact elements of tooth flanks are refined (with size 0.05 mm) to capture the nature of tooth contact. Usually,
2D surface to surface contact is applied (see [46] for details). Rigid couplings are established between the nodes at inner bores and
the corresponding centers. The torque 𝑇 is applied to the pinion, while the rotation is applied to the gear. Starting at the initial
mesh position 𝛼 = 𝛽1 , two mesh cycle are simulated. Within the mesh duration, 200 load steps are included. In each load step, the
gear rotation is uniformly increased, and the data including rotated angle of rigid coupling, and contact forces in each iteration
are extracted by Python® scripts [53]. Focusing on the simulation of mesh stiffness rather than deformed tooth profiles, unlike the
previous work [46], the mesh stiffness is calculated through the torque–rotation method: 𝐾m = 𝑇 ∕𝑅𝑏,p ∕(𝜃p 𝑅𝑏,p + 𝜃g 𝑅𝑏,g ).
The load sharing is a precondition for the evaluation of mesh stiffness, which also dictates the mesh mechanisms and the
subsequent dynamics. The predicted load sharing ratio 𝜌𝑙 = 𝑅𝑏,p 𝐹𝑙 ∕𝑇 (𝑙 = 1, 2, 3 represents the tooth pair 𝑙) of gear pair A (plane
strain) are compared the FEM in Fig. 5.2. Good correspondences between the present model (continuous lines) and simulations
(dashed lines) are observed for both gear pairs.
Due to the extended tooth contact (ETC), the theoretically separated tooth will engage. As a result, smooth transitions between
STC and DTC are reflected in both the present model and the FEM. On the other hand, as explicitly represented in Eq. (2.22), the
extended tooth contact is load-dependent. Compared with the smaller torque (𝑇 = 25 N m, Fig. 5.2(a)), the larger torque (𝑇 =
50 N m, Fig. 5.2(b)) further decreases the STC region RSTC and increases the contact ratio 𝜀(𝑇 ). A more detailed discussion on this
phenomenon will be performed in Section 6.1
As emphasized before, the present model aims to cover both thick-walled gears and thin-walled gears. Similar comparisons for
gear pair B (thin-walled gear, plane stress) are shown in Fig. 5.3. Again, the predicated result (continuous lines) matches well
with the FEM (dashed lines). The load dependency of the contact ratio 𝜀(𝑇 ) is also observed. Comparisons in Figs. 5.2 and 5.3
demonstrated the ability of the president model to cover both thick-walled gears and thin-walled gears.
The corresponding mesh stiffness for gear pair A and gear pair B are further compared in Figs. 5.4 and 5.5, respectively. In
order to check the robustness under different value of the ratio ℎ = 𝑅𝑓 ∕𝑅int , wide ranges of inner bore radius for each gear pair are
covered : 𝑅int = 4, 8, 12 mm (i.e. ℎ ∈ [1.56, 4.69]) for gear pair A ; 𝑅int = 10, 30, 50 mm (i.e. ℎ ∈ [1.43, 7.13]) for gear pair B.
For a more comprehensive contrast, the polynomial formulas of Xie et al. [26] is also included in the comparison (Figs. 5.4 and
5.5, gray dashed lines). In the mesh stiffness model of Xie et al. [26], the polynomial formulas is controlled by 198 non-dimensional
constants with confidence intervals ℎ ∈ [2, 5), 𝜃𝑓 ∈ [0.04, 0.08] in plane strain. The load sharing is further calculated through the
minimum potential energy method [54] based on theoretical intervals of DTC and STC.

12
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Fig. 5.2. Load sharing ratio 𝜌𝑙=1,2,3 calculated by the present model (continuous lines with: 𝜌1 —, 𝜌2 —, and 𝜌3 —) and FEM (dashed lines with: 𝜌1 - -, 𝜌2 - -,
and 𝜌3 - -) for Gear pair A (plane strain, ICR = 1.71, 𝑅int = 8 mm) under different torques: (a) 𝑇 = 25 N m; (b) 𝑇 = 50 N m.

Fig. 5.3. Load sharing ratio 𝜌𝑙=1,2,3 calculated by the present model (continuous lines with: 𝜌1 —, 𝜌2 —, and 𝜌3 —) and FEM (dashed lines with: 𝜌1 - -, 𝜌2 - -,
and 𝜌3 - -) for Gear pair B (plane stress, ICR = 1.75, 𝑅int = 30 mm) under different torques: (a) 𝑇 = 50 N m; (b) 𝑇 = 100 N m.

For thick-walled gears (gear pair A), observation of Fig. 5.4 indicates close mesh stiffness curves between the present model (red
lines) and FEM (black lines). Compared with the polynomial formulas [26], the present model exhibits satisfactory robustness for
the change of inner bore radius. Beyond the confidence intervals of polynomial formulas, ℎ ∈ [2, 5), the present model still matches
well with the FEM (see Fig. 5.4(c), ℎ = 1.56). It can be observed that the stiffness arises with the increase of hub bore radius, since
the reduced elastic region of fillet foundation becomes hard to deform. On the other hand, as already discussed for the results of
load-sharing, the extended tooth contact (ETC) reduces the interval of STC. This phenomenon is also reflected in the mesh stiffness
curves. Both FEM (black lines) and the present model (red lines) exhibit smooth transitions between STC and DTC, while traditional
method presents abrupt changes (gray dashed lines).
The present model also covers thin-walled gears, which are widely employed in high-speed transmissions for their light-
weighting. Analogously, mesh stiffness curves for gear pair B are further compared in Fig. 5.5. For thin-walled gears, it can be
observed that the accuracy of the present model remains: the predicated results (Fig. 5.5(a)–(c), blue lines) again coincide well with
the FEM (Fig. 5.5(a)–(c), black lines).
A series of comparisons of load sharing ratio (Figs. 5.2 and 5.3) and mesh stiffness (Figs. 5.4 and 5.5) demonstrated the reasonable
accuracy of the present model. The wider application range and higher accuracy in the prediction of deformation responses lay
foundations for the subsequent dynamic analysis.

5.2. Dynamic responses

Not satisfied verifications only in quasi-statics, this section further extended the comparison in the domain of dynamics.
Experimental data from A. Kahraman et al. [49,52] are taken as benchmarks. The experiment of A. Kahraman covered a wide
range of mesh frequency (500 Hz √ < 𝑓m < 3500 Hz) for gear pair B. The equivalent root-mean-square (rms) of oscillating DTE
∑3 2
component is measured by: 𝐴rms = 𝑛=1 𝐴𝑛 (𝐴𝑛 is the 𝑛−th harmonic amplitude of DTE).

13
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Fig. 5.4. Mesh stiffness calculated by the present model (—), FEM (—), and polynomial formulas [26] (- -) for gear pair A (plane strain, ICR = 1.71, 𝑇 = 25 N
m) with different inner bore radius: (a) 𝑅int = 4 mm; (b) 𝑅int = 8 mm; (c) 𝑅int = 12 mm. (For interpretation of the references to color in this figure legend, the
reader is referred to the web version of this article.)

Fig. 5.5. Mesh stiffness calculated by the present model (—), FEM (—), and polynomial formulas [26] (- -) for gear pair B (plane stress, ICR = 1.75, 𝑇 = 100 N
m) with different inner bore radius: (a) 𝑅int = 10 mm; (b) 𝑅int = 30 mm; (c) 𝑅int = 50 mm. (For interpretation of the references to color in this figure legend,
the reader is referred to the web version of this article.)

Fig. 5.6. 𝐴rms of gear pair B (unmodified, 𝑅int = 22.5 mm, 𝜁 = 0.04) reproduced by the present model (stepping up (⋅), stepping down (◦)) and experiments (□):
(a) ICR = 1.37, 𝑇 =170 N m, natural frequency 𝑓𝑛 = 2740 Hz; (b) ICR = 1.77, 𝑇 =170 N m, natural frequency 𝑓𝑛 = 2900 Hz.

With the stepping up and stepping down process, the predicated 𝐴rms of gear pair B are compared to experiments in Fig. 5.6. Due
to the uncertainty of manufacturing and measurement, discrepancies still exist between the present model and experimental data.
Nevertheless, close trends are reflected in both amplitudes and resonances with ICR = 1.37 (Fig. 5.6(a)) and ICR = 1.77 (Fig. 5.6(b)),
which demonstrated the effectiveness of the present model.
Subsequently, to emphasize the sensitivity of mesh stiffness in the prediction of dynamic responses, different mesh stiffness
waveforms considering ETC, ignoring ETC, and using polynomial formulas are compared in Fig. 5.7. Due to the flexibility of gear
pair, the stiffness curves considering ETC (Fig. 5.7(a), red curves) presents larger contact ratio compared with the theoretical contact
regions (Fig. 5.7(a), blue and gray curves). According to Eq. (2.24), the contact ratio becomes 𝜀(𝑇 ) = 1.58 in the case of considering
ETC, while other cases remain the standard contact ratio, namely 𝜀∗ = 1.37.

14
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Fig. 5.7. Mesh stiffness waveforms of gear pair B (ICR = 1.37, 𝑅int = 22.5 mm, 𝑇 = 170 N m) obtained by polynomial formulas (—), the present model ignoring
ETC (—), and the present model considering ETC (—): (a) mesh stiffness curves; (b) 4-order Fourier representations. (For interpretation of the references to
color in this figure legend, the reader is referred to the web version of this article.)

Fig. 5.8. Dynamic responses of gear pair B (ICR = 1.37, 𝑅int = 22.5 mm, 𝑇 = 170 N m) using different mesh stiffness waveforms (Fourier representation): (a)
overview; (b) enlarged view. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

As indicated in Eq. (3.3), the extended contact ratio 𝜀(𝑇 ) influences the representation of mesh stiffness, which positively
correlates to the average component 𝐾0 and dictates harmonic contents 𝐾𝑛 . As a consequence, the 4-order Fourier representations of
the corresponding mesh stiffness curves in Fig. 5.7(b) present distinct magnitudes as well as tendencies. In sense of dynamics, these
changing harmonic contents (i.e. 𝐾0 , 𝐾𝑛 ) further excite different resonances of gear pairs, for example super-harmonic resonances
and primary resonance.
The 4-order Fourier waveforms in Fig. 5.7(b) are further used as inputs to the dynamic equation in Eq. (3.7). Dynamic responses
with polynomial formulas (gray circles), the present model ignoring ETC (blue circles), and the present model ignoring ETC (red
circles) are compared in Fig. 5.8. It can be observed that all cases present similar trends compared to the experimental data.
However, great differences appear at the primary and super-harmonic resonance regions due to the different representations of
stiffness harmonic contents. As clearly reflected in the enlarged view (Fig. 5.8(b)), the case of considering ETC are closer to the
experiment compared with other cases.
Next, Fig. 5.9 further compared the dynamic response (𝐴rms ) calculated by the present model (considering ETC) with Chen
et al. [55] (ignoring ETC) and experimental results [52]. Similar to the present work, Chen’s model also refined the mesh stiffness
by considering the structure coupling effect in simultaneous tooth contact. As shown in Fig. 5.9(a), both the present model (red
circles) and Chen’s model (blue triangles) exhibit close jump frequencies and vibration amplitudes to experimental data (black
squares). Whereas, due to the ETC effect considered by this model, an evident difference arises in the super-harmonic resonance
regions (Fig. 5.9(b)). As discussed before, stiffness harmonic contents altered by ETC will influence the dynamic responses. As a
result, the present model shows a conspicuous 𝑓𝑛 ∕4 resonance which is also reflected in the experimental data. Both Figs. 5.8 and
5.9 demonstrated that considering ETC could reasonably capture the experimental resonances and amplitudes. In gear dynamic
modeling, the incorporation of extended tooth contact (ETC) is recommended.
From quasi-statics (Section 5.1) to dynamics (Section 5.2), the reasonable accuracy of the present model is verified through a
series comparisons with specially-designed FEM, existing models, and experiments. Comparisons in Sections 5.1–5.2 also reflect the

15
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Fig. 5.9. 𝐴rms of gear pair B (ICR = 1.75, 𝑅int = 22.5 mm, 𝑇 = 340 N m) reproduced by the present model (◦), digitized from Chen, et al. (▵), and experiment
(□): (a) overview; (b) enlarged view.

Fig. 6.1. Load sharing ratio 𝜌2 of different gear pairs under uniformly increasing torques: (a) gear pair A (ICR = 1.71, 𝑅int = 8 mm) with torque 𝑇 = 10–50 N
m; (b) gear pair B (ICR = 1.75, 𝑅int = 30 mm) with torque 𝑇 = 50–200 N m.

significant influences of the extended tooth contact on both deformation and dynamic responses. Of particular interest in finding
some regularities on this phenomenon, more detailed parametric studies will be performed in Section 6.

6. Parametric studies

6.1. Effects of extended tooth contact on deformation responses

This section attempts to investigate the effect of extended tooth contact on the deformation responses in depth. Similar to
Section 5.1, both gear pair A and gear pair B are included in the present study.
Fig. 6.1 displaced the load sharing ratio 𝜌2 (referring to tooth pair 2) with uniformly increasing torques for of gear pair A
(Fig. 6.1(a)) and gear pair B (Fig. 6.1(b)). In the double tooth contact region RDTC , curves of 𝜌2 slightly changes in terms of slopes,
since the stiffness component 𝐾ℎ is load-dependent. As indicated by Chen [32], if the term is replaced with 𝐾ℎ = 𝜋𝐸𝐿∕4(1 − 𝜈)2 , the
slope variation vanishes. Unlike DTC, the load-dependency of the load sharing ratio 𝜌2 in extended tooth contact (ETC) is clearly
reflected by both gear pair A (Fig. 6.1(a)) and B (Fig. 6.1(b)). Due to the gear pair flexibilities, the theoretically separated teeth are
forced to contact. As a result, slopes of 𝜌2 curves have to become sharp to maintain the deformation compatibility.
Differing to the profile modifications, the extended tooth contact stems from the gear flexibility and deformation compatibility,
which is nonlinearly influenced by the applied torque. An overview on the influence of torque is summarized in Fig. 6.2(a)(b). The
region of single tooth contact (STC) non-uniformly narrows with increasing torques, which leads to nonlinear increment of regions
for simultaneous tooth contact.
Both Figs. 6.1–6.2 indicate the nonlinear influences of torque on the tooth contact. As introduced in Section 2.2.5, the nonlinear
increments of regions in simultaneous tooth contact is quantified with the term 𝜀(𝑇 ) (so-called extended contact ratio, ECR). Fig. 6.3
clearly reflected the load-dependency of the extended tooth contact ratio. It can be observed that the ECR gradually and nonlinearly
arises with increasing torques, while the ICR maintains the same level.

16
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Fig. 6.2. Contour maps of load sharing ratio 𝜌2 for different gear pairs: (a) gear pair A (ICR = 1.71, 𝑅int = 8 mm) with torque 𝑇 = 10–50 N m; (b) gear pair
B (ICR = 1.75, 𝑅int = 30 mm) with torque 𝑇 = 50–200 N m.

Fig. 6.3. Involute contact ratio (ICR) and extended tooth contact ratio (ECR): (a) gear pair A (ICR = 1.71, 𝑅int = 8 mm) with torque 𝑇 = 10–50 N m; (b) gear
pair B (ICR = 1.75, 𝑅int = 30 mm) with torque 𝑇 = 50–200 N m.

The contact ratio increased by torque will further influence the mesh stiffness. The corresponding mesh stiffness curves and
4-th order Fourier waveforms for gear pair A and gear pair B are shown in Fig. 6.4(a)–(d). Results for both gear pairs indicate the
sensitivity of mesh stiffness to the applied torque.
The load-dependency of mesh stiffness mainly derives from two aspects. As clearly indicated in Fig. 6.4(a)(c), the magnitude of
stiffness curves increases with the torque, since the load–deflection relation of tooth pair is nonlinear depending on the nonlinear
contact between tooth interfaces and the structural coupling of fillet foundation. On the other hand, as indicated by Fig. 6.3,
the increasing torque enlarges the contact ratio 𝜀(𝑇 ). As a result, both mesh stiffness curves (see Fig. 6.4(a)(c)) and Fourier
representations (see Fig. 6.4(b)(d)) subsequently change with the varying tooth contact modes, namely STC, DTC, and ETC.
Fig. 6.5 (gear pair A) and Fig. 6.6 (gear pair B) further compared 4-orders harmonic contents (namely, 𝐾𝑛 , 𝑛 = 1, 2, 3, 4) in the
case of considering ETC and ignoring ETC. Harmonic contents in the case of ignoring ETC (see Figs. 6.5(b) and 6.6(b)) very slowly
grow with the increasing torque, which is mainly caused by the load-dependency of contact stiffness and the structural coupling of
fillet foundation. Interestingly, in the case of considering ETC (see Figs. 6.5(a) and 6.6(a)), all orders of harmonic contents sensitively
vary with the increasing torque.
Results in Figs. 6.5 and 6.6 demonstrated that harmonic contents 𝐾𝑛 are strongly influenced by the torque. Compared with the
other factors (e.g. nonlinear tooth contact, structural coupling of fillet foundation), 𝐾𝑛 appear more sensitive to the extended tooth
contact. Due to the nonlinear relation between the torque 𝑇 and the contact ratio 𝜀(𝑇 ), the varying spectra of mesh stiffness would
further induce substantial differences in the dynamic responses.

6.2. Effects of extended tooth contact on dynamic responses

As concluded in Section 6.1, due to the extended tooth contact (ETC), the load-dependent contact ratio 𝜀(𝑇 ) dictates the harmonic
contents 𝐾𝑛 , playing a predominant role on the representation mesh stiffness. By controlling variables, the proposed mesh stiffness

17
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Fig. 6.4. Mesh stiffness under increasing torques: (a)(b) stiffness curves and Fourier representations of gear pair A (ICR = 1.71, 𝑅int = 8 mm, 𝑇 = 10–50 N m);
(c)(b) stiffness curves and Fourier representations of gear pair B (ICR = 1.75, 𝑅int = 30 mm, 𝑇 = 50–200 N m).

Fig. 6.5. Fourier amplitudes of mesh stiffness for gear pair A (ICR = 1.71, 𝑅int = 8 mm, 𝑇 = 10–50 N m): (a) considering ETC; (b) ignoring ETC.

waveforms in both considering and ignoring ETC are used in the dynamic model. Consequently, the effect of ETC on dynamic
responses can be clearly reflected.
Similar to Section 5.2, the gear pair B (ICR = 1.37, 𝑅int = 22.5 mm, natural frequency 𝑓𝑛 = 2740 Hz) are adopted. In the present
analysis, the range of applied torque is 𝑇 = 20–200 N m, and the mesh frequency ranges 𝑓m = 500–3500 Hz. For clarity, the
corresponding Fourier contents of mesh stiffness are also summarized in Fig. 6.7, which outlines the subsequent dynamic responses.

Fig. 6.8 compared the 𝐴rms in cases of considering ETC (lines with points) and ignoring ETC (dashed lines). Both results in
stepping up process (Fig. 6.8(a)) and stepping down process (Fig. 6.8(b)) indicate that the vibration intensifies with the increasing
torque 𝑇 due to the rising energy input of the system. In each case, curves of 𝐴rms present similar bifurcations near the primary
resonance. Frequency shifts can be also observed in the stepping back (Fig. 6.8(b)), since the average stiffness average mesh stiffness

18
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Fig. 6.6. Fourier amplitudes of mesh stiffness for gear pair B (ICR = 1.75, 𝑅int = 30 mm, 𝑇 = 50–200 N m): (a) considering ETC; (b) ignoring ETC.

Fig. 6.7. Comparison of Fourier contents in cases of considering ETC (continuous lines with markers) and ignoring ETC (dashed lines with markers) for gear
pair B (ICR = 1.37, 𝑅int = 22.5 mm, 𝑇 = 20–200 N m): (a) average mesh stiffness value 𝐾0 ; (b) harmonic contents 𝐾𝑛 , 𝑛 = 1, 2, … , 4.

Fig. 6.8. Comparison of 𝐴rms in cases of considering ETC (lines with points) and ignoring ETC (dashed lines) for gear pair B (ICR = 1.37, 𝜁 = 0.04, 𝑇 =
25–200 N m): (a) stepping up, 𝑓m = 500 → 3500 Hz; (b) stepping down, 𝑓m = 3500 → 500 Hz.

value 𝐾0 is directly correlated to the contact ratio (see Fig. 6.7(a)). Apart from average mesh stiffness value 𝐾0 , the extended contact
ratio also dictates the harmonic contents 𝐾𝑛 , which lead to substantial differences in the dynamic responses (e.g. superharmonic
and its following resonances). As clearly reflected in Fig. 6.8(a)(b), within the region 500 Hz–1600 Hz, distinct differences of the
multiple super-harmonic resonances are observed between the cases of considering and ignoring ETC.

19
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Fig. 6.9. Contour maps of multiple super-harmonic resonances in terms of 𝐴rms near 500 Hz–1600 Hz: (a) with ETC; (b) without ETC.

More intuitively, the variation of resonances are further visualized through the contour maps in Fig. 6.9. Cases of considering and
ignoring ETC present distinct distributions of multiple super-harmonic resonances (𝑓𝑛 /2, 𝑓𝑛 /3, and 𝑓𝑛 /4). In Fig. 6.9(b) (ignoring
ETC), the 𝑓𝑛 /2 super-harmonic resonance is predominant. Interestingly, if the extended tooth contact is considered (see Fig. 6.9(a)),
the variation trends of 𝐾𝑛 completely changes. As a consequence, the predominance of 𝑓𝑛 /2 super-harmonic resonance is weakened,
while the sensitivity of 𝑓𝑛 /3 super-harmonic resonance arises. In fact, as discussed in Fig. 5.8 (see Section 5.2), similar distributions
of multiple super-harmonic resonances are also found in the experiment: the 𝑓𝑛 /2 super-harmonic resonance is more noticeable than
𝑓𝑛 /3.
Combining the Fourier contents in Fig. 6.7(a)(b), the complicated phenomenon in Fig. 6.9 can be qualitatively explained from
spectra. Due to the higher levels of the average value 𝐾0 and the fundamental harmonic 𝐾1 , the 𝑓𝑛 ∕2 superharmonic resonance
in the case ETC presents lower amplitude and higher frequency compared with the case without ETC. Meanwhile, the nonlinear
variations of 𝐾𝑛 in terms of 𝑇 together change the multiple resonances.
The sensitivity of super-harmonic resonances are also reflected in the time domain. By tracking the resonance peaks in the
stepping up process, steady-state solutions of phase portraits at the primary resonance (𝑓m = 𝑓𝑛 ) and multiple super-harmonic
resonances (𝑓m = 𝑓𝑛 ∕2, 𝑓𝑛 ∕3, and 𝑓𝑛 ∕4) are further compared in Fig. 6.10, which comprehensively summarized the influences of
ETC on the gear vibration and stableness under the varying torque 𝑇 . With the increasing torque, the intensifying energy further
expands the orbits of 𝑞m − 𝑞̇ m and forms the ‘‘tornado’’ shapes of phase portraits in Fig. 6.10(a)–(d). Here 𝑞̇ m = 𝑑𝑞𝑚 ∕𝑑𝑡 represents
the vibration velocity.
For the primary resonance (Fig. 6.10(a)), due to the change of mean value 𝐾0 (refer to Fig. 6.7(a)), it can be observed that the
variation trend as well as amplitude of phase portraits are slightly perturbed. Compared with the primary resonance, oscillations
in multiple super-harmonic resonances are more sensitive to the extended tooth contact. As shown in Fig. 6.10(c)–(d), noticeable
differences can be observed between dashed lines (ignoring ETC) and solid lines (considering ETC). The extended tooth contact
greatly alters the Fourier representations of mesh stiffness, leading to diversities of dynamic responses.
From Sections 6.1–6.2, effects of extended tooth contact (ETC) on both deformation responses and dynamic responses are
quantitatively analyzed. Superficially, from the perspective of mesh stiffness curves, no essential differences exist between the
stiffness curves (see e.g. Fig. 5.7(a)): the extended tooth contact (ETC) only creates additional smooth transitions between single
tooth contact (STC) and double tooth contact (DTC), which maybe one of the reason why many analytical models ignored the
extended tooth contact.
However, these inconspicuous smooth transitions further influence the load sharing and enlarge the contact ratio 𝜀(𝑇 ). As a
consequence, mesh stiffness curves present distinct Fourier representations. In the case of considering ETC, both the average value
𝐾0 and harmonic contents 𝐾𝑛 are nonlinearly correlated with the torque 𝑇 , which further perturb the subsequent primary resonance
and super-harmonic resonances.
In the analysis of gear pair B with ICR = 1.37, the extended tooth contact leads to higher peak value of 𝑓𝑛 ∕2 resonance than
𝑓𝑛 ∕3 resonance (see Fig. 6.9(a)(b)). Interestingly, similar results can also be observed in the experiment [49] and some FEM-based
results [18,46]. From this perspective, the rationality of mesh stiffness in ETC is demonstrated to some extent. On the other hand,
for spur gears, the extended tooth contact is a natural reaction stemming from the elasticity of tooth pairs and fillet foundation.
Additional consideration on the extended tooth contact facilitates a deeper understanding of the optimization designs and failure
analysis.

7. Conclusion

From kinematics to dynamics, this paper developed an analytical approach to bridge the extended tooth contact and gear
nonlinear dynamics. A series of comparisons with FEM, theoretical models, and experiments demonstrated the reasonable accuracy
of the present model on both quasi-statics and dynamics. Findings and contributions of this paper are summarized as follows:

20
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Fig. 6.10. Steady-state solutions of phase portraits in cases of considering ETC (solid lines) and ignoring ETC (dashed lines): (a) the primary resonance, 𝑓𝑚 = 𝑓𝑛 ;
(b) 𝑓𝑛 ∕2 super-harmonic resonance; (c) 𝑓𝑛 ∕3 super-harmonic resonance; and (d) 𝑓𝑛 ∕4 super-harmonic resonance.

• The present model extended the bi-dimensional formulas of fillet foundation for a wider application range of modern gears.
Both thick-walled gears and thin-walled gears are covered, while restrictions of ℎ and 𝜃𝑓 are removed. Compared with the
widely used polynomial formulas [25,26], the predicated mesh stiffness curves are closer to FEM results, while the robustness
and efficiency are remained.
• Based on small deformation assumption, an analytical representation on the extended tooth contact region is proposed,
which avoided iterations on distinguishing tooth contact modes and kept the acceptable correspondences with FEM. With
the proposed mesh stiffness waveform, the nonlinear dynamic model is further extended to embody the load-dependency of
extended tooth contact.
• Parametric studies on the extended tooth contact is performed. The investigation shows that extend tooth contact dictates
the Fourier representation of mesh stiffness waveforms. The alterant harmonic contents will further influence behaviors of
multiple super-harmonic resonances. For the mesh stiffness modeling, considerations on the extended tooth contact and actual
contact ratio are recommended.

Focusing on the correlation between extended tooth contact and dynamics for a basic transmission element (a pair of spur
gear), the multi-DOF dynamics, the tooth profile modification, and the extreme case (triple tooth contact) are temporarily excluded.
Nevertheless, the proposed model could be extended as an interface to the subsequent dynamic modeling, failure analysis, and
optimization designs for multi-DOF systems.

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 National Science and Technology Major Project (No:J2019-IV-0001-0068).

21
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Appendix A. Expressions of 𝑪𝒏 (𝜿, 𝒉) and 𝑫𝒏 (𝜿, 𝒉)

Expressions of 𝐶𝑛 (𝜅, ℎ) and 𝐷𝑛 (𝜅, ℎ) in Eq. (2.11) write:


(𝜅 + 1)(ℎ2 − 1)ℎ2
𝐶𝑛 (𝜅, ℎ) = , (A.1)
(ℎ2 − 1)2 (1 − 𝑛2 ) − (ℎ2𝑛+2
+ 𝜅)(ℎ−2𝑛+2 + 𝜅)
⎧ 𝜅(ℎ2𝑛+2 −1)(ℎ−2𝑛+2 +𝜅)
⎪ 1+𝑛
+ (1 − 𝑛)(ℎ2 − 1)2
⎪ − , 𝑛 ≠ −1
2 2 2 2𝑛+2 + 𝜅)(ℎ−2𝑛+2 + 𝜅)
𝐷𝑛 (𝜅, ℎ) = ⎨ (ℎ − 1) (1 − 𝑛 ) − (ℎ . (A.2)
⎪ 2(ℎ2 − 1)2 2𝜅
⎪ + ln ℎ, 𝑛 = −1
⎩ (𝜅 + 1)(𝜅 + ℎ4 ) 𝜅 + 1
During the numerical computation, fractional splitting forms of 𝐶𝑛 (𝜅, ℎ) and 𝐷𝑛 (𝜅, ℎ) (𝑛 ≠ −1) write:
⎧ (𝜅 + 1)(ℎ2 − 1)ℎ2
⎪𝐶 ∗ (𝜅, ℎ) =
⎪ 𝑛 (ℎ2 − 1)2 (1 − 𝑛2 ) − ℎ4 − 𝜅(ℎ2𝑛+2 + ℎ−2𝑛+2 ) − 𝜅 2
⎨ , (A.3)
⎪𝐷∗ (𝜅, ℎ) = − 𝜅∕(𝑛 + 1) + (1 − 𝑛)(ℎ2 − 1)2 ∕(ℎ4 + 𝜅ℎ2𝑛+2 − ℎ−2𝑛+2 − 𝜅)
⎪ 𝑛 (ℎ2 − 1)2 (1 − 𝑛2 )∕(ℎ4 + 𝜅ℎ2𝑛+2 − ℎ−2𝑛+2 − 𝜅) − 1 − (𝜅 + 1)(ℎ2𝑛+2 − 1)

which efficiently removed the numerical problem ‘‘inf ∗ 0 = nan’’ for the computation of infinite terms.

Appendix B. Evaluations of 𝜳𝒔 , 𝜣𝒔𝟐,𝟏 , and 𝜣𝒔𝟏,𝟐

The dimensionless coefficients 𝛹𝑠 , (𝑠 = 0, 1, 2) and 𝛩𝑠𝑎,𝑏 , (𝑠 = 0, 1, 2, 3) hold:


[ ] [ ]
1+𝜈 1 ∑ ∞
1+𝜈 1 ∑∞
𝛹𝑠 = − 𝐺𝑠 (0) + 𝐺𝑠 (𝑛) , 𝛩𝑠𝑎,𝑏 = − 𝐻𝑠𝑎,𝑏 (0, 𝑘) + 𝐻𝑠𝑎,𝑏 (𝑛, 𝑘) . (B.1)
2𝜋 2 𝑛=0
2𝜋 2 𝑛=0

Here, 𝐺𝑠 (𝑛) and 𝐻𝑠𝑎,𝑏 (𝑛, 𝑘) are the 𝑛−th items in infinite series 𝛹𝑠 and 𝛩𝑠𝑎,𝑏 , respectively. Subscripts ∙𝑎,𝑏 represent the labels of teeth
(𝑎, 𝑏 = 1, 2, and 𝑎 ≠ 𝑏); The integer 𝑘 = ±1 is an indicator as defined in Fig. 2.1.
Expressions of items 𝐺𝑠 (𝑛), (𝑠 = 0, 1, 2) are given as follows:
⎧𝐺 (𝑛) = 𝛿 𝛥(1) − 𝛾 𝛯 (2) + 𝜆 𝛬 tan2 𝛼
⎪ 0 𝑛 𝑛
[
𝑛 𝑛 𝑛 𝑛 𝐹
]
⎪ ( ) (1) 𝛿𝑛 (2) 𝛾𝑛 (1) ( ) (2)
⎨𝐺1 (𝑛) = 2𝜃𝑓 𝛿𝑛 + 𝑛𝜆𝑛 𝛥𝑛 + 2𝜃 𝛥𝑛 − 2𝜃 𝛯𝑛 + 𝑛𝛿𝑛 − 𝛾𝑛 𝛯𝑛 . (B.2)
⎪ 𝑓 𝑓
⎪𝐺 (𝑛) = 2𝜃 [(𝑛𝛿 − 𝛾 ) 𝛯 (1) + (𝑛𝜆 + 𝛿 ) 𝛥(2) ]
⎩ 2 𝑓 𝑛 𝑛 𝑛 𝑛 𝑛 𝑛

Expressions of items 𝐻𝑠𝑎,𝑏 (𝑛, 𝑘), (𝑠 = 0, 1, 2, 3) are given as follows:


( ) ( )
⎧ 𝑎,𝑏 2𝜋𝑛𝑘 tan 𝛼𝐹 𝑏 2𝜋𝑛𝑘 tan 𝛼𝐹 𝑏
⎪𝐻0 (𝑛, 𝑘) = cos ⋅ 𝛿𝑛 𝛥(1) 𝑛 − 𝛾𝑛 𝛯𝑛(2) + 𝜆𝑛 𝛬𝑛 + sin ⋅ tan 𝛼𝐹 𝑎 𝜆𝑛 𝛥(1)
𝑛 + 𝛿𝑛 𝛯𝑛(2) − 𝛿𝑛 𝛬𝑛
⎪ 𝑧 cot 𝛼𝐹 𝑎 𝑧 tan 𝛼𝐹 𝑎
⎪ 𝑎,𝑏 2𝜋𝑛𝑘 ( (2) (1)
) 2𝜋𝑛𝑘 ( (2) (1)
)
⎪𝐻1 (𝑛, 𝑘) = cos 𝑧 ⋅ 𝛿𝑛 𝛥𝑛 − 𝛾𝑛 𝛯𝑛 + sin 𝑧 ⋅ tan 𝛼𝐹 𝑎 𝜆𝑛 𝛥𝑛 + 𝛿𝑛 𝛯𝑛
⎨ , (B.3)
⎪𝐻 𝑎,𝑏 (𝑛, 𝑘) = 2𝜃 cos 2𝜋𝑛𝑘 [(𝛿 + 𝑛𝜆 )𝛥(1) + (𝑛𝛿 − 𝛾 ) 𝛯 (2) ] − 2𝜃 sin 2𝜋𝑛𝑘 ⋅ tan 𝛼 (𝛿 + 𝑛𝜆 )𝛬
⎪ 2 𝑓
𝑧 𝑛 𝑛 𝑛 𝑛 𝑛 𝑛 𝑓
𝑧 𝐹𝑏 𝑛 𝑛 𝑛
⎪ 𝑎,𝑏 2𝜋𝑛𝑘 (2) (1)
⎪𝐻3 (𝑛, 𝑘) = 2𝜃𝑓 cos ⋅ [(𝛿𝑛 + 𝑛𝜆𝑛 )𝛥𝑛 + (𝑛𝛿𝑛 − 𝛾𝑛 )𝛯𝑛 ]
⎩ 𝑧
where 𝜆𝑛 , 𝛿𝑛 , and 𝛾𝑛 are linear combinations of 𝐶±𝑛 (𝜅, ℎ) and 𝐷±𝑛 (𝜅, ℎ):
𝜆𝑛 = 𝐶𝑛 (𝜅, ℎ) + 𝐶−𝑛 (𝜅, ℎ) + 𝐷𝑛 (𝜅, ℎ) + 𝐷−𝑛 (𝜅, ℎ)
𝛿𝑛 = 𝐶𝑛 (𝜅, ℎ) − 𝐶−𝑛 (𝜅, ℎ) + 𝐷𝑛 (𝜅, ℎ) − 𝐷−𝑛 (𝜅, ℎ) (B.4)
𝛾𝑛 = 𝐶𝑛 (𝜅, ℎ) + 𝐶−𝑛 (𝜅, ℎ) − 𝐷𝑛 (𝜅, ℎ) − 𝐷−𝑛 (𝜅, ℎ)
In Eqs. (B.2)–(B.3), 𝛬𝑛 , 𝛯𝑛 , and 𝛥𝑛 are integrals and can be obtained according to [26]:
⎧ sin(𝑛𝜃𝑓 )
⎪ ,𝑛≠0
⎪ 𝑛 sin 𝜃𝑓 𝑢 𝑆𝑓 (1)
𝛬𝑛 = ⎨ , 𝛯𝑛 = 𝐹 𝛯𝑛(1) + 𝛯𝑛(2) , 𝛥𝑛 = 𝛥 + 𝛥(2)
𝑛 . (B.5)
⎪ 𝜃 𝑓 𝑆 𝑓 𝑢𝐹 𝑛
,𝑛=0
⎪ sin 𝜃𝑓

𝛯𝑛(1) and 𝛯𝑛(2) are components of 𝛯𝑛 , expressed as:
[ ]
⎧ sin(𝑛𝜃𝑓 ) 2 cos(𝑛𝜃𝑓 ) 2 sin(𝑛𝜃𝑓 ) ⎧ sin(𝑛𝜃𝑓 )
⎪ 6 + − ,𝑛≠0 ⎪ ,𝑛≠0
𝛯𝑛(1) = ⎨ 𝑛 𝑛2 𝜃𝑓 𝑛3 𝜃𝑓2 ; 𝛯𝑛(2) = ⎨ 𝑛𝜃𝑓 . (B.6)
⎪ ⎪1, 𝑛 = 0
⎩2𝜃𝑓 , 𝑛 = 0 ⎩

22
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Fig. D.1. Separation distance of the pinion when the gear is fixed.

𝛥(1) (2)
𝑛 , and 𝛥𝑛 are components of 𝛥𝑛 , written as:

⎧ 𝜃𝑓 (sin 𝜃𝑓 − 𝜃𝑓 ) ⋅ 𝜃 3 sin(𝑛𝜃)
⎪𝛥(1) = 1 𝑑𝜃
⎪ 𝑛 2𝜃𝑓 ∫−𝜃𝑓 3𝜃 2 sin 𝜃𝑓 + 6𝜃𝑓 cos 𝜃𝑓 − 𝜃 3 cos 𝜃𝑓 − 6 sin 𝜃𝑓
⎪ (
𝑓 𝑓
)
⎨ −1 3
(B.7)
⎪ (2) 𝜃𝑓 6 𝜃𝑓 sin 𝜃𝑓 + 2 cos 𝜃𝑓 − 2𝜃𝑓 sin 𝜃𝑓 ⋅ 𝜃 sin(𝑛𝜃)
1
⎪𝛥𝑛 = 𝑑𝜃
⎪ 2𝜃𝑓 ∫−𝜃𝑓 3𝜃 2 sin 𝜃𝑓 + 6𝜃𝑓 cos 𝜃𝑓 − 𝜃 3 cos 𝜃𝑓 − 6 sin 𝜃𝑓
⎩ 𝑓 𝑓

Appendix C. Angular positions for theoretical STC and DTC

According to the involute geometry, expressions of 𝛽1 , 𝛽2 , 𝛽3 , and 𝛽4 are given by:


( )
𝜋 𝑅𝑏,g 𝑅𝑏,g 2𝜋
𝛽1 = 𝛼n − , 𝛽2 = 1 + tan 𝛼n − tan 𝛼𝑎,g − 𝜃𝑏,p + ,
2𝑧p 𝑅𝑏,p 𝑅𝑏,p 𝑧𝑝
(C.1)
𝜋 3𝜋
𝛽3 = 𝛼𝑎,p − + inv(𝛼𝑎,p ) − inv(𝛼n ), 𝛽4 = 𝛼n + ,
2𝑧p 2𝑧p
where inv(𝑥) = tan 𝑥 − 𝑥 is the involute function; 𝛼n is the pressure angle; 𝛼𝑎,p and 𝛼𝑎,g are the pressure angles at tip circles of pinion
and gear, respectively. 𝑧p is the tooth number of pinion.

Appendix D. The tooth separation function 𝝓(𝜶)

For clarity, the tooth separation distance evaluation in Ref. [48] is illustrated in Fig. D.1. As stated before, the present model
regarded the pinion as the driving side and the gear as the driven side. As a consequence, in the modeling of each quasi-static time
frame, the pinion is moving while the gear can be treated as fixed.
Starting from a generic mesh location 𝛼, the pinion tooth rotates from the point 𝑁0 to the theoretical engaging-in point 𝑁2 .
During the rotation, the pinion tooth collides with the corresponding gear tooth at the point 𝑀, as the gear is fixed. As a result,
concyclic points 𝑁0 , 𝑁1 , and 𝑁2 formed the trajectory of the approaching pinion tooth (see the locus circle in Fig. D.1). The so-called
tooth separation function 𝜙(𝛼) thus refers to the angle ∠𝑁0 𝑂p 𝑁1 = 𝜙 with varying initial mesh location 𝛼.
In light of Ref. [48], geometric quantities and spatial relationship of the approaching tooth pair are labeled in Fig. D.1. Here,
𝜓p and 𝜓g are rolling angles of the pinion and the gear, which are the only variables in the equations in Fig. D.1. The connections
between 𝜓p , 𝜓g and the input 𝛼 can be readily established by: 𝜓p (𝛼) = 𝛽2 − 2𝜋∕𝑧p − 𝛼 and 𝜓g (𝛼) = 𝜓p (𝛼)𝑧p ∕𝑧g . Finally, the tooth
separation function can be evaluated with 𝜙(𝛼) = 𝜓p − 𝜃1 + 𝜃2 + 𝜃3 .

23
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

Table E.1
Geometrical and material properties of spur gear pairs.
Properties Gear pair A Gear pair B
Gear Pinion Gear Pinion
Tooth number 𝑧 40 40 50 50
Modules 𝑚 (mm) 1 1 3 3
Tooth width 𝐿 (mm) 20 20 20 20
Pressure angle 𝛼n (◦ ) 20 20 20 20
Young’s modulus (GPa) 206.8 206.8 206.8 206.8
Poisson’s ratio 0.3 0.3 0.3 0.3
Density (kg/m3 ) 7850 7850 7850 7850
Hub radius 𝑅int (mm) 4–12 10–50
Tip clearance coefficient 0.25 0.25
Addendum coefficient 1.00 0.76 1.00 1.01
Involute contact ratio 1.71 1.37 1.75 1.77

Appendix E. Properties of specimens

See Table E.1.

References

[1] J. Wang, R. Li and, X. Peng, Survey of nonlinear vibration of gear transmission systems, Appl. Mech. Rev. 56 (3) (2003) 309–329, http://dx.doi.org/10.
1115/1.1555660.
[2] M. Khabou, N. Bouchaala, F. Chaari, T. Fakhfakh, M. Haddar, Study of a spur gear dynamic behavior in transient regime, Mech. Syst. Signal Process. 25
(8) (2011) 3089–3101, http://dx.doi.org/10.1016/j.ymssp.2011.04.018.
[3] S. Theodossiades, S. Natsiavas, Non-linear dynamics of gear-pair systems with periodic stiffness and backlash, J. Sound Vib. 229 (2) (2000) 287–310,
http://dx.doi.org/10.1006/jsvi.1999.2490.
[4] S. Natsiavas, S. Theodossiades, I. Goudas, Dynamic analysis of piecewise linear oscillators with time periodic coefficients, Int. J. Nonlinear Mech. 35 (1)
(2000) 53–68, http://dx.doi.org/10.1016/S0020-7462(98)00087-0.
[5] J. Wang, R. Li, X. Peng, Survey of nonlinear vibration of gear transmission systems, Appl. Mech. Rev. 56 (3) (2003) 309–329, http://dx.doi.org/10.1115/
1.1555660.
[6] A. Kahraman, G. Blankenship, Interactions between commensurate parametric and forcing excitations in a system with clearance, J. Sound Vib. 194 (3)
(1996) 317–336, http://dx.doi.org/10.1006/jsvi.1996.0361.
[7] R.G. Munro1, D. Palmer, L. Morrish, An experimental method to measure gear tooth stiffness throughout and beyond the path of contact, Proc. Inst. Mech.
Eng. C 215 (7) (2001) 793–803, http://dx.doi.org/10.1243/0954406011524153.
[8] 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, http://dx.doi.org/10.1016/j.engfailanal.2013.07.005.
[9] N.K. Raghuwanshi, A. Parey, Mesh stiffness measurement of cracked spur gear by photoelasticity technique, Measurement 73 (2015) 439–452, http:
//dx.doi.org/10.1016/j.measurement.2015.05.035.
[10] C. Natali, M. Battarra, G. Dalpiaz, E. Mucchi, A critical review on FE-based methods for mesh stiffness estimation in spur gears, Mech. Mach. Theory 161
(2021) 104319, http://dx.doi.org/10.1016/j.mechmachtheory.2021.104319.
[11] J. Wang, I. Howard, Finite element analysis of high contact ratio spur gears in mesh, J. Tribol. 127 (3) (2005) 469–483, http://dx.doi.org/10.1115/1.
1843154.
[12] N.L. Pedersen, M.F. Jørgensen, On gear tooth stiffness evaluation, Comput. Struct. 135 (2014) 109–117, http://dx.doi.org/10.1016/j.compstruc.2014.01.023.
[13] Z. He, Y. Hu, X. Zheng, Y. Yu, A calculation method for tooth wear depth based on the finite element method that considers the dynamic mesh force,
Machines 10 (2) (2022) 69, http://dx.doi.org/10.3390/machines10020069.
[14] J. Zhan, M. Fard, R. Jazar, A CAD-FEM-QSA integration technique for determining the time-varying meshing stiffness of gear pairs, Measurement 100
(2017) 139–149, http://dx.doi.org/10.1016/j.measurement.2016.12.056.
[15] X. Liang, H. Zhang, M.J. Zuo, Y. Qin, Three new models for evaluation of standard involute spur gear mesh stiffness, Mech. Syst. Signal Process. 101
(2018) 424–434, http://dx.doi.org/10.1016/j.ymssp.2017.09.005.
[16] K. Chen, Y. Huangfu, H. Ma, Z. Xu, X. Li, B. Wen, Calculation of mesh stiffness of spur gears considering complex foundation types and crack propagation
paths, Mech. Syst. Signal Process. 130 (2019) 273–292, http://dx.doi.org/10.1016/j.ymssp.2019.05.014.
[17] R. Parker, S. Vijayakar, T. Imajo, Non-linear dynamic response of a spur gear pair: Modelling and experimental comparisons, J. Sound Vib. 237 (3) (2000)
435–455, http://dx.doi.org/10.1006/jsvi.2000.3067.
[18] C.G. Cooley, C. Liu, X. Dai, R.G. Parker, Gear tooth mesh stiffness: A comparison of calculation approaches, Mech. Mach. Theory 105 (2016) 540–553,
http://dx.doi.org/10.1016/j.mechmachtheory.2016.07.021.
[19] J.D. Marafona, P.M. Marques, R.C. Martins, J.H. Seabra, Mesh stiffness models for cylindrical gears: A detailed review, Mech. Mach. Theory 166 (2021)
104472, http://dx.doi.org/10.1016/j.mechmachtheory.2021.104472.
[20] C. Weber, The deformAtion of Loaded Gears and the Effect on Their Load-Carrying Capacity, Report 3, Sponsored Research (Germany), British Dept. of
Scientific and Industrial Research, 1949.
[21] D.C.H. Yang, J.Y. Lin, Hertzian damping, tooth friction and bending elasticity in gear impact dynamics, J. Mech. Transm. Autom. Des. 109 (2) (1987)
189–196, http://dx.doi.org/10.1115/1.3267437.
[22] X. Tian, Dynamic Simulation for System Response of Gearbox including Localized Gear Faults (Master’s Thesis), University of Alberta, Edmonton, Alberta,
Canada, 2004.
[23] A. Palmgren, Ball and Roller Bearing Engineering, SKF Industries Inc, Philadelphia, 1959.
[24] Muskhelishvili, Some Basic Problems of the Mathematical Theory of Elasticity, Noordhoff Groningen, 1953.
[25] P. Sainsot, P. Velex, O. Duverger, Contribution of gear body to tooth deflections-a new bidimensional analytical formula, J. Mech. Des. 126 (4) (2004)
748–752, http://dx.doi.org/10.1115/1.1758252.

24
X. Zheng et al. Mechanism and Machine Theory 175 (2022) 104958

[26] C. Xie, L. Hua, X. Han, J. Lan, X. Wan, X. Xiong, Analytical formulas for gear body-induced tooth deflections of spur gears considering structure coupling
effect, Int. J. Mech. Sci. 148 (2018) 174–190, http://dx.doi.org/10.1016/j.ijmecsci.2018.08.022.
[27] 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,
http://dx.doi.org/10.1016/j.engfailanal.2012.08.015.
[28] H. Ma, R. Song, X. Pang, B. Wen, Time-varying mesh stiffness calculation of cracked spur gears, Eng. Fail. Anal. 44 (2014) 179–194, http://dx.doi.org/
10.1016/j.engfailanal.2014.05.018.
[29] Z. Chen, J. Zhang, W. Zhai, Y. Wang, J. Liu, Improved analytical methods for calculation of gear tooth fillet-foundation stiffness with tooth root crack,
Eng. Fail. Anal. 82 (2017) 72–81, http://dx.doi.org/10.1016/j.engfailanal.2017.08.028.
[30] Z. Chen, W. Zhai, K. Wang, Vibration feature evolution of locomotive with tooth root crack propagation of gear transmission system, Mech. Syst. Signal
Process. 115 (2019) 29–44, http://dx.doi.org/10.1016/j.ymssp.2018.05.038.
[31] 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, http://dx.doi.org/10.1016/j.mechmachtheory.2015.11.017.
[32] Z. Chen, Z. Zhou, W. Zhai, K. Wang, Improved analytical calculation model of spur gear mesh excitations with tooth profile deviations, Mech. Mach.
Theory 149 (2020) 103838, http://dx.doi.org/10.1016/j.mechmachtheory.2020.103838.
[33] F. Bruzzone, T. Maggi, C. Marcellini, C. Rosso, 2D nonlinear and non-hertzian gear teeth deflection model for static transmission error calculation, Mech.
Mach. Theory 166 (2021) 104471, http://dx.doi.org/10.1016/j.mechmachtheory.2021.104471.
[34] 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 transmissions,
Mech. Mach. Theory 61 (2013) 30–58, http://dx.doi.org/10.1016/j.mechmachtheory.2012.10.008.
[35] J.I. Pedrero, M. Pleguezuelos, M.B. Sánchez, Influence of meshing stiffness on load distribution between planets of planetary gear drives, Mech. Mach.
Theory 170 (2022) 104718, http://dx.doi.org/10.1016/j.mechmachtheory.2021.104718.
[36] M. Pleguezuelos, M.B. Sánchez, J.I. Pedrero, Analytical model for meshing stiffness, load sharing, and transmission error for spur gears with profile
modification under non-nominal load conditions, Appl. Math. Model. 97 (2021) 344–365, http://dx.doi.org/10.1016/j.apm.2021.03.051.
[37] H. Nevzat Özgüven, D. Houser, Dynamic analysis of high speed gears by using loaded static transmission error, J. Sound Vib. 125 (1) (1988) 71–83,
http://dx.doi.org/10.1016/0022-460X(88)90416-6.
[38] T. Eritenel, R.G. Parker, Nonlinear vibration of gears with tooth surface modifications, J. Vib. Acoust. 135 (5) (2013) http://dx.doi.org/10.1115/1.4023913,
051005.
[39] V.K. Tamminana, A. Kahraman, S. Vijayakar, A study of the relationship between the dynamic factors and the dynamic transmission error of spur gear
pairs, J. Mech. Des. 129 (1) (2006) 75–84, http://dx.doi.org/10.1115/1.2359470.
[40] P. Velex, M. Maatar, A mathematical model for analyzing the influence of shape deviations and mounting errors on gear dynamic behaviour, J. Sound
Vib. 191 (5) (1996) 629–660, http://dx.doi.org/10.1006/jsvi.1996.0148.
[41] X. Gu, P. Velex, A dynamic model to study the influence of planet position errors in planetary gears, J. Sound Vib. 331 (20) (2012) 4554–4574,
http://dx.doi.org/10.1016/j.jsv.2012.05.007.
[42] Z. Cao, Z. Chen, H. Jiang, Nonlinear dynamics of a spur gear pair with force-dependent mesh stiffness, Nonlinear Dynam. 99 (2) (2020) 1227–1241,
http://dx.doi.org/10.1007/s11071-019-05348-0.
[43] H. Lin, J. Wang, F. Oswald, J. Coy, Effect of extended tooth contact on the modeling of spur gear transmissions, in: 29th Joint Propulsion Conference
and Exhibit, 1993, p. 2148.
[44] C. Xie, X. Shu, A new mesh stiffness model for modified spur gears with coupling tooth and body flexibility effects, Appl. Math. Model. 91 (2021)
1194–1210, http://dx.doi.org/10.1016/j.apm.2020.11.003.
[45] R.W. Cornell, Compliance and stress sensitivity of spur gear teeth, J. Mech. Des. 103 (2) (1981) 447–459, http://dx.doi.org/10.1115/1.3254939.
[46] X. Zheng, W. Luo, Y. Hu, Z. He, S. Wang, Study on the mesh stiffness and nonlinear dynamics accounting for centrifugal effect of high-speed spur gears,
Mech. Mach. Theory 170 (2022) 104686, http://dx.doi.org/10.1016/j.mechmachtheory.2021.104686.
[47] 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, http://dx.doi.org/10.1016/j.mechmachtheory.2012.10.012.
[48] D. Tse, H. Lin, Separation distance and static transmission error of involute spur gears, in: 28th Joint Propulsion Conference and Exhibit, 1992, p. 3490.
[49] A. Kahraman, G.W. Blankenship, Effect of involute contact ratio on spur gear dynamics, J. Mech. Des. 121 (1) (1999) 112–118, http://dx.doi.org/10.
1115/1.2829411.
[50] H. Nevzat Özgüven, D. Houser, Mathematical models used in gear dynamics - a review, J. Sound Vib. 121 (3) (1988) 383–411, http://dx.doi.org/10.1016/
S0022-460X(88)80365-1.
[51] A. Kahraman, R. Singh, Non-linear dynamics of a spur gear pair, J. Sound Vib. 142 (1) (1990) 49–75, http://dx.doi.org/10.1016/0022-460X(90)90582-K.
[52] A. Kahraman, G.W. Blankenship, Experiments on nonlinear dynamic behavior of an oscillator with clearance and periodically time-varying parameters, J.
Appl. Mech. 64 (1) (1997) 217–226, http://dx.doi.org/10.1115/1.2787276.
[53] X. Zheng, S. Wang, Spur gear extraction, 2022, http://dx.doi.org/10.13140/RG.2.2.19666.48321.
[54] 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, http://dx.doi.org/10.1016/j.ymssp.2018.03.037.
[55] Z. Chen, J. Ning, K. Wang, W. Zhai, An improved dynamic model of spur gear transmission considering coupling effect between gear neighboring teeth,
Nonlinear Dynam. 106 (1) (2021) 339–357, http://dx.doi.org/10.1007/s11071-021-06852-y.

25

You might also like