You are on page 1of 22

Mechanical Systems and Signal Processing 173 (2022) 109052

Contents lists available at ScienceDirect

Mechanical Systems and Signal Processing


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

Torsional vibration responses of the engine crankshaft-gearbox


coupled system with misfire and breathing slant crack based on
instantaneous angular speed
Kun Wu a, Zhiwei Liu a, Qian Ding a, b, *, Fengshou Gu c, Andrew Ball c
a
Department of Mechanics, Tianjin University, Tianjin 300350, China
b
Tianjin Key Laboratory of Nonlinear Dynamics and Control, Tianjin 300350, China
c
School of Computing and Engineering, University of Huddersfield, Huddersfield HD1 3DH, UK

A R T I C L E I N F O A B S T R A C T

Communicated by Didier Remond Torsional vibration responses of an engine crankshaft-gearbox coupled system with misfire and
breathing slant crack are investigated in this paper. The crankshaft is modelled by lumped mass
Keywords: method and its dynamic equations are deduced based on Lagrange equation. Both the time
Torsional vibration varying mesh stiffness of the gearbox composing of a pinion and a gear and the response-
Crankshaft
dependent stiffness of a shaft element with a breathing slant crack are evaluated. Natural fre­
Gearbox
quencies of the healthy and cracked systems are analysed respectively. Torsional vibrations based
Misfire
Crack on the instantaneous angular speed (IAS) of the flywheel and the driven shaft are compared
Instantaneous angular speed among the operating conditions of normal, misfire and breathing slant crack on the driven shaft.
Simulation results are validated by experiments, which show that the encoder on the driven shaft
is better because it can monitor the IAS of the flywheel and the driven shaft simultaneously and
the signals are more prominent here.

1. Introduction

Diesel engine is the primary driver in plenty of industrial applications, including compressors, generators and on and off road
vehicles [1]. As one of the most important components, crankshaft of the diesel engine is subjected to complex external and internal
excitations [2]. The resulted torsional vibrations are the most dangerous for the crankshaft [3,4]. Therefore, it is crucial to detect and
diagnose the conditions of the crankshaft through instantaneous angular speed (IAS) to prevent early faults and damages like misfire
and breathing slant crack. However, the install space of monitoring sensors might be limited for certain diesel engines because of the
timing belt/chain on the free end of the crankshaft, and the flywheel connected to the gearbox directly. Thus, it is important to study
the coupled torsional vibration responses of the engine crankshaft-gearbox system if the monitoring sensors are installed on the output
end of the gearbox.
Potenza et al. built [5] a two-degree-of freedom dynamical model of a multi-cylinder engine coupled to a dynamometer to simulate
the instantaneous crank kinematics, and the model could predict the total mechanical losses accurately at high-speed. Chilinski and
Zawisza [6] analysed the modelling of crank-piston mechanism with a torsional vibration damper and validated it with experiments.
They found that the coupling between the equations of the bending and angular oscillations is strong. Karabulut [7] pointed out that

* Corresponding author at: Department of Mechanics, Tianjin University, Tianjin 300350, China.
E-mail address: qding@tju.edu.cn (Q. Ding).

https://doi.org/10.1016/j.ymssp.2022.109052
Received 20 June 2021; Received in revised form 16 February 2022; Accepted 11 March 2022
Available online 24 March 2022
0888-3270/© 2022 Elsevier Ltd. All rights reserved.
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

the engine vibrations are originated from inertial effects of piston-crankshaft mechanism and gas forces, then the author investigated
the periodic and temporary variations of crankshaft speed and the variation of the torque and power with respect to the damping and
stiffness coefficients of the mounts. Ni et al. [8] investigated the effect of advanced injection angle on shaft torsional vibrations based
on a coupling model of shaft using Simulink and concluded that the frequency of cylinder pressure fluctuations might affect the
torsional vibration responses. Liu et al. [9] studied the dynamic responses of an engine-coupling-dynamometer system based on the
torque and the torsional vibration of the elastic coupling. Based on investigation of the coupled vibrations of a propeller-crankshaft-
sliding bearing system in longitudinal and torsional directions, Zhang et al. [10] pointed out that the strength calculation and vibration
analysis of the system should be based on nonlinear model to achieve high accuracy when designing the propeller shafts used in marine
diesel engines. Pfabe and Woernle [11] developed a kinematical driven flywheel to compensate the fluctuating engine output torque
and reduce the torsional vibrations, and tested it on an electrically driven stand. By modelling a six-cylinder four-stroke inline diesel
engine’s crankshaft using AFL Excite Designer, Sun et al. [12] found that the moment of inertia contributed by the timing gear could be
ignored. Li et al. [13] considered the friction force of the piston pack ring as one of the excitation sources to obtain the torsional
vibration of the crankshaft, and the results showed that the variation in torsional amplitude at 2.0 order frequency could be the sign of
scuffing failure of the diesel engine.
The gear transmission is one of the most important systems for motion and power transmission [14], and the time-varying meshing
stiffness of the gear pair will cause nonlinear dynamic responses [15]. Using short-time Fourier transform, Li and Pang [16] analysed
the single crack, gear coupling crack, single shaft crack, and gear and shaft coupling crack signals, respectively, based on finite element
method. They found that the shaft and gear crack might bring in the side frequency offsets in FFT spectrum. Chaari et al. [17] deduced
an analytical formulation of the time varying gear mesh stiffness of the gear with cracked tooth, and it is in good accordance with the
finite element model. Sainsot and Velex [18] presented an improved fillet/foundation compliance analysis based on the theory of
Muskhelishvili, and obtained the analytical formula of gear body-induced tooth deflections. Chen and Shao proposed an analytical
model to study the effect of gear tooth crack on the gear mesh stiffness, and came up with the conclusion that the sidebands caused by
the tooth crack are more sensitive than the mesh frequency and its harmonics in [19]. Another analytical model took gear tooth
deviation into consideration by Chen and Shao in [20], which proved to be suitable for the gear pairs with both high and low contact
ratio. A gear mesh kinematic model was built by Luo et al. [21], which calculated the actual contact position of tooth engagement
based on the time varying gear mesh center distance. Chen et al. [22] conducted numerical simulations to evaluate the effects of engine
torque fluctuations and tooth surface friction on the gear rattle responses and the corresponding tooth impact behaviour fluctuation,
and showed that they mainly affect the high frequency components. Luo et al. [23] evaluated the time varying mesh stiffness with a
curved-bottom shaped tooth spall, which corrected the foundation stiffness within the double tooth contact area. Two key factors,
namely the boundary condition through the gear rim size and the size of the contact, were considered in [24] when the time varying
mesh stiffness was calculated. Yu et al. [25] indicated that spatial crack propagation in the crack depth direction, the tooth width
direction and the tooth profile direction would reduce the gear mesh stiffness and load sharing ratio between the gear pair. Ma et al.
[26] improved the analytical model for the healthy gear pair by treating the gear tooth as the nonuniform beam on the root circle, and
considering the misalignment of gear root circle and base circle and accurate transition curve.
Misfire of combustion is a typical and serious problem a diesel engine often meets, which can directly lead to working condition
deterioration, additional vibration and noise, and then reduce the stability and reliability of the engine [27–30]. Through the iden­
tification of the generalized force at the engine center of gravity, Xu et al. [31] proposed a method on detecting the misfire fault, in
which the acceleration signals at the mounts were analysed by discrete spectrum interpolation method. Jung et al. [32] presented an
algorithm for misfire detection to compensate the vehicle-to-vehicle variations based on the crankshaft angular velocity measured at
the flywheel, and indicated that the flywheel error adaption will improve robustness and reduce the number of mis-classifications. By
experiments, the authors also validated that this misfire detection algorithm only has to be calibrated using data from one vehicle. A
novel method to detect cylinder misfire through sound quality metrics of the radiated sound was proposed by Singh et al. [33], aiming
at avoiding direct contact with hot vibrating engine component(s). Considering some quality metrics including loudness, roughness
and fluctuation strength of engine and tailpipe sounds, the support vector machine proved to be 94 % accuracy in experiments. Chen
and Randall [34] used artificial neural networks to diagnose accurately the location and severity of misfire of an internal combustion
engine. Cavina et al. [35] came up with a methodology by pre-processing the combustion time intervals to increase the signal-to-noise
ratio. By monitoring the responses of each cylinder, the methodology would correct the combustion time signal time every cycle. Tests
show that the threshold values and the correction vector are quite robust. Helm et al. [36] used two Kalman filters for each cylinder to
represent the normal and the misfire condition, and this approach can achieve early and reliable detection of the actual cylinder state.
Sharma et al. [37] compared J48 algorithm, best first tree algorithm, random forest tree algorithm, functional tree algorithm and linear
model tree algorithm based on the statistical features extracted from engine vibration signals. The decision tree algorithms are easier to
implement as an onboard system.
Catastrophic failures of rotating machinery may start with small defects and cracks on the shaft, thus it is important to study the
vibrations of cracked rotor and provide theoretical and experimental basis for detection and online monitoring [38]. Xiang et al. [39]
established a nonlinear model, which took the rub-impact forces, the time varying stiffness of cracked shaft and the nonlinear oil-film
forces into account. The dynamic responses of rotor were reflected by orbits, bifurcation diagrams, time series and frequency spectra.
Han and Chu [40] studied the lateral vibrations of the transverse-surface cracked rotor with an elliptical front, and found that the
instability regions increase towards lower rotating speed area with deeper surface crack. Based on compliance method and strain
energy release rate method, they calculated the stiffness coefficients of the cracked rotor with rotationally asymmetric inertia in [41],
and the unstable widths of the Jeffcott rotor were obtained through the harmonic balance method and Taylor expansion technique.
Ramezanpour et al. [42] evaluated the flexibility matrix and then the stiffness matrix of a Jeffcott rotor system with slant crack

2
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

orienting in arbitrary directions, and concluded that the slant crack would lead the dynamic responses to differences between forward
and backward whirls. A new model for a crankshaft with a slant crack in crankpin was proposed by Lei et al. [43]. Simulation results
show that the crack could cause prominent changes in the spectra of transient responses. Effects of a straight crack on the coupled
vibrations in longitudinal, lateral and torsional directions were discussed by Darpe et al. in [44], which indicates that the breathing
crack is dependent on the responses. When studying the coupling mechanisms between the rotor’s vibrations in translational and
rotational degrees of freedoms, Zhang and Li [45] calculated the crack closure line in each step, and used 200 statuses to describe the
open and close conditions of the crack during rotation. Numerical results show that a perturbation frequency component and its
combinations with harmonic frequencies would appear in the dynamic responses.
To the best of the authors’ information, the coupled torsional vibration of the engine crankshaft-gearbox system with misfire and
slant crack cases remains unsolved. In this paper, the dynamic equations of the engine crankshaft are established in Section 2. Then, the
time varying mesh stiffness of the gearbox is analysed and added to the above equations, which is followed by the stiffness calculation
of the breathing slant crack on the output shaft of the gearbox. After that, numerical simulations are carried in Section 3. At last, the
results are validated by experiments in Section 4.

2. Dynamic modelling of the engine crankshaft-gearbox coupled system

2.1. The engine crankshaft-gearbox coupled system

The investigated system includes two-cylinder Beta 14 diesel engine and a TMC40P gearbox, as shown in Fig. 1. A torsional vi­
bration absorber, two cylinders and a flywheel are arranged from left to right inside the engine. In the gearbox, a spur pinion is
connected with the flywheel by a clutch, while a spur gear is connected with the elastic coupling for power output. The driven shaft
transmits the torque and rotational movement from the gearbox to propeller in the tank full of circulating water.
Schematic of the 8 degree-of-freedom engine crankshaft-gearbox coupled system (the coupled system hereinafter) is shown in
Fig. 2. The subscripts 0 to 3 represent the torsional vibration absorber, the first cylinder, the second cylinder and the flywheel of engine
crankshaft, respectively. The pinion and gear are distinguished by the subscripts p and g. Subscript c represents the elastic coupling. A
45-degree breathing slant crack on the surface of the driven shaft (subscript driven) will be considered for the crack case. Note that the
clutch between the flywheel and the pinion remains engaged in this paper. The corresponding parameters of the crankshaft are listed in
Table. 1.

2.2. Modelling of the engine crankshaft

The non-constant inertia model of the crankshaft can be built based on the lumped mass method [9,46]. To establish the analytical
dynamic equation of the crankshaft, we first derive the non-constant instantaneous inertia of single cylinder, and then deduce the
dynamic equation based on Lagrange equation.

Fig. 1. The test rig of the engine crankshaft-gearbox coupled system.

3
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

Fig. 2. Structural schematic of the engine crankshaft-gearbox coupled system.

Table 1
Structural parameters of the crankshaft.
Parameter Value

I0 1.81 × 105 kg∙mm2


I3 3 × 105 kg∙mm2
K0 5.15 × 105 N∙mm/rad
K1 2.055 × 107 N∙mm/rad
K2 2.055 × 107 N∙mm/rad
K3 2.03 × 107 N∙mm/rad
C1 100 N∙mm∙s/rad
C2 100 N∙mm∙s/rad
ω0 10.2127 Hz

2.2.1. Non-constant inertia of single cylinder


As shown in Fig. 3, for a single cylinder, the piston reciprocates from left to right, the connecting rod swings and the crank rotates
around its axis during operation. Physical meanings and values of the parameters in the figure are listed in Table 2.
From the geometrical relationships of the crank-connecting rod mechanism, the total instantaneous kinetic energy is determined as
1[ ] 2 1 2 1( )
Ek = Icrank + Iwi + m1 r2 + mc (hr)2 θ̇ + Ia β̇ + m2 + mp ẋ2p . (1)
2 2 2
where

Ia = Ir − m1 (jL)2 − m2 (L − jL)2
m1 = mr (1 − j) , (2)
m2 = jmr
and Ir is the inertia of the connecting rod.
Using the instantaneous kinetic energy equivalence method, the following equation can be obtained,

Fig. 3. Structural schematic of a single cylinder.

4
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

Table 2
Structural parameters of the cylinder.
Parameter Physical meaning Value

Icrank Inertia of the crank 1.78 × 104 kg∙mm2


Iwi Inertia of the counterweight 1 × 104 kg∙mm2
mr Mass of the connecting rod 1.14 kg
mp Mass of the piston 1.2078 kg
mc Mass of the crank 2.6478 kg
L Length of the connecting rod 118 mm
r Radius of the crank 34 mm
h Length ratio of OA to OG 0.1039
j Length ratio of AG’ to AB 0.29
Dc Bore 67 mm
G Mass center of the crank –
G’ Mass center of the connecting rod –

1
(3)
2
Ek = IE (θ)θ̇ ,
2
where IE(θ) is the equivalent inertia of a single cylinder.
According to the geometric structure in Fig. 2, the relationships rsinθ = Lsinβ and xp = rcosθ + Lcosβ can be obtained, where xp is
( )
the displacement of the piston in x axis. Thus, one has β̇ = rcosθ rsin2θ
Lcosβ and ẋp = − rθ̇ sinθ + 2Lcosβ . Substitute them into Eq. (1) and combine
θ̇

with Eq. (3), the equivalent inertia can be expressed as


( )
IE (θ) = Icrank + Iwi + mc (hr)2 + (1 − j)mr r2 + jmr + mp r2 sin2 θ. (4)

Let
1( )
Ie = Icrank + Iwi + mc (hr)2 + (1 − j)mr r2 + jmr + mp r2
2
( ) , (5)
jmr + mp r2
ε=
2Ie
Eq. (4) can be simplified as IE(θ) = Ie(1 − εcos2θ).

2.2.2. Dynamic equation of the engine crankshaft


Dynamic equation of the engine crankshaft in torsional direction will be deduced based on Lagrange equation as
⎛ ⎞
d ⎝ ∂T ⎠ ∂T ∂U ∂D
− + + = Mi , (6)
dt ∂θ̇i ∂θi ∂θi ∂θ̇i

where i = 0, 1, 2 or 3, and Mi is the external torque applied to the i-th lumped mass. For the torsional vibration absorber (i = 0), the
external torque is zero. For each cylinder (i = 1 or 2), it mainly originates from the gas pressure, see discussion in Section 2.2.3. For the
flywheel (i = 3), it is the load T3 = K3(φp-φ3) from the pinion.
In Eq. (6) the kinetic energy is determined by

1 2 1 2 1∑ 2
(7)
2
T = I0 θ̇0 + I3 θ̇3 + IE (θi )θ̇i ,
2 2 2 i=1

the potential energy is

1∑ 3
U= Ki (φi − φi+1 )2 , (8)
2 i=0

and the dissipative energy is


( )2
1∑ 2
1∑ 2
(9)
2
D= Ci θ̇i + Ci,i+1 φ̇i − φ̇i+1
2 i=1 2 i=0

respectively. θi is angular displacement of the corresponding node and φi is its torsional vibration with relation
θ i = ω t + φ i + ξi , (10)
where ω is rotating speed, t the rotating time, and ξi values 0, 0, π and 0, respectively.
In Eqs. (7)–(9), Ii, Ki and Ci denote inertia, torsional stiffness and damping respectively. The internal damping of shaft is determined

5
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

by Ci,i+1 = 0.02Ki/πω0, where ω0 is the first natural frequency of the crankshaft in torsional direction [46].

2.2.3. External torque of the cylinder


The external torque of each cylinder (i = 1 or 2) consists of the torque due to the non-constant inertia Mnc, the reciprocating inertia
torque Mre and the combustion gas torque Mp, where
1′
Mnc = − ωC − I (θ)ω2 , (11)
2E

( ) ( r ) sin(θ + β)
Mre = − jmr + mp r2 ω2 cosθ + cos2θ , (12)
L cosβ

πD2c sin(θ + β)
Mp = rPg . (13)
4 cosβ
Mnc is the additional terms in Eq. (6) due to the non-constant inertia of the cylinder, I’E(θ) the derivative of IE(θ), and Pg the in-
cylinder gas pressure obtained by experiments.
The time waveforms and spectra of the external torque of two cylinders are shown in Fig. 4 in case the rotating frequency f = 20 Hz
(note the rotating speed ω = 2πf). Because of the firing order and the structural characteristics, there is a phase gap of π between two
cylinders in the time waveform. It is worth to point out that the linear combinations of the harmonics of θ, 2θ and θ + β in Eqs (12) and
(13) will result that the frequency components at 0.5, 1, 1.5, 2.5 times of f are dominant.
Note that the mean value of external torque of each cylinder is the driving torque, and it should be eliminated when the torsional
vibration of the coupled system is calculated.

2.3. Time varying mesh stiffness of the gearbox

Torque and rotational movement are transmitted by the gearbox through the elastic deformations of the gear pairs in engagement
[47], which can be considered as a nonlinear spring along the line of action with time varying stiffness. The mesh force between the
pinion and gear is calculated as
( )
Fmesh = Km φp Rp + φg Rg , (14)

where Km is the time varying mesh stiffness, φp/g the torsional vibration of the pinion/gear and Rp/g the radius of the pitch circle of the
pinion/gear. Km varies periodically because both the number of the tooth pairs in mesh and the contact position along the tooth profile
vary periodically.
For a meshed tooth pair, mesh stiffness Ksingle is obtained based on potential energy method [23,26].
1 1 1 1 1 1 1 1 1 1
= + + + + + + + + , (15)
Ksingle Kh Kap Kag Kbp Kbg Ksp Ksg Kfp Kfg

where Kh, Ka, Kb, Ks and Kf are Hertzian contact, axial compressive, bending, shear and fillet foundation stiffness respectively. The
corresponding stiffness of the pinion and the gear are distinguished with the subscripts p and g. Table 3 lists structural parameters of the

Fig. 4. (a) Time waveforms and (b) spectra of the external torque of two cylinders.

6
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

gearbox.
Supposing that there is no variation of the center distance between the pinion and the gear, the total time varying mesh stiffness is
estimated by

N
Km = Ksingle (i), (16)
i=1

where N is number of the meshed tooth pairs at present time (N = 1 or 2 in this paper). The time varying mesh stiffness of the
investigated pinion-gear pair is shown in Fig. 5.

2.4. Stiffness of the driven shaft with slant crack

Periodical excitation may induce a crack on the shaft of the rotating machinery during long-term operation. Appearance of 45-de­
gree slant crack on shaft surface is the most frequently occurred faults in rotating machines under torsional excitations. We consider a
45-degree slant crack on the driven shaft surface, which will close and open due to the torsional vibration difference between the
connected shafts. The schematic of the driven shaft and the cross section of the crack are shown in Fig. 6.
Addition torsional displacement due to a crack with the depth of α can be determined by [43]
[∫ α ]

μ= J(α)dα , (17)
∂Tdriven 0
where Tdriven is the torque transmitted by the driven shaft and J the strain energy density function.
1[ ]
J= (KI )2 + m(KIII )2 , (18)
E′
where E’ =E for plane stress problem, E’ =E/(1-ν2) for plane strain problem, m = 1+µ, and KI and KШ are the crack intensity factors,
2Tdriven x √̅̅̅̅̅̅
KI = παFII (s)
π R4
√̅̅̅ √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ , (19)
2 2Tdriven R2 − x2 √̅̅̅̅̅̅
KIII = πα F III (s)
π R4

where
√̅̅̅̅̅̅̅̅̅̅̅̅̅
2
1.122− 0.561s+0.85s +0.18s3 2 πs
s = √̅̅√α̅̅̅̅̅̅̅̅̅̅,
2 2 R2 − x2
FII = √̅̅̅̅̅̅
1− s
,FIII = πstan 2 .
The flexibility of the driven shaft without crack μ0 = l/GmIdriven, where l is the length of the shaft element, and Gm is the shear
modulus. Thus, the torsional stiffness Kdriven = 1/μ when the sign of φdriven-φc is positive, namely the crack is closed. Kdriven = 1/(μ + μ0)
in other situations, namely the crack is open. φc and φdriven are the torsional vibration of the coupling and the driven shaft, respectively.
Corresponding parameters of the driven shaft are listed in Table 4.

2.5. Dynamic equations of the coupled system

Based on the Newton’s second law of motion, the dynamic equations of the pinion and the gear are deduced as
Ip φ̈p = − T3 + Fmesh Rp
, (20)
Ig φ̈g = Fmesh Rg + Tc

where Tc = Kc(φc − φg) is the driven torque from the coupling.


Similarly, the dynamic equations of the coupling and the driven shaft are
Ic φ̈c = − Tc + Tdriven
, (21)
Idriven φ̈driven = − Tdriven + Tload
where Tdriven = Kdriven(φdriven-φc) and Tload is the varying part of the external load from the immersed rotating propeller. Because of
the gearbox, the rotating frequency of the propeller fdriven = fNp/Ng. Considering that the torque of this three-blade propeller is the
function of its geometric parameters, the external load of the driven shaft [48] is

Table 3
Structural parameters of the gearbox.
Parameter Physical meaning Value

mp/g Module of the pinion and the gear 2 mm


αp/g Pressure angle at the pitch circle 20◦
Np Tooth number of the pinion 22
Ng Tooth number of the gear 47

7
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

Fig. 5. Time varying mesh stiffness.

Fig. 6. (a) The driven shaft with breathing slant crack and (b) the cross section of the crack.

Table 4
Structural parameters of the driven shaft.
Parameter Physical meaning Value

E Young’s modulus 2.1 × 1011 Pa


ν Poisson’s ratio 0.3
α Depth of the crack 3 mm
l Length of the driven shaft 150 mm
R Radius of the driven shaft 15 mm

Tload = {A1 sin(2πfdriven t) + A2 sin[2(2πfdriven t) ] + A3 sin[3(2π fdriven t) ] }, (22)


where A1, A2 and A3 are the parameters related to the pitch distance of the propeller.
Combine Eq. (6) with Eq. (21) and Eq. (22), dynamic equation of the coupled system can be obtain as
Mφ̈ + Cφ̇ + Kφ = T, (23)

where M = diag[I0 , IE (θ1 ), IE (θ2 ), I3 , Ip , Ig , Ic , Idriven ],

8
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

Fig. 7. Simulation procedure.

9
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

⎡ ⎤
⎢ K0 − K0 0 0 0 0 0 0 ⎥
⎢ ⎥
⎢ ⎥
⎢ − K0 K0 + K1 − K1 0 0 0 0 0 ⎥
⎢ ⎥
⎢ 0 − K1 K1 + K2 − K2 0 0 0 0 ⎥
⎢ ⎥
⎢ ⎥
⎢ 0 0 − K2 K2 + K3 − K3 0 0 0 ⎥
K=⎢
⎢ 0
⎥,
⎢ 0 0 − K3 K3 + Km R2p K m Rp Rg 0 0 ⎥ ⎥
⎢ ⎥
⎢ 0 0 0 0 K m Rp Rg Km R2g + Kc − Kc 0 ⎥
⎢ ⎥
⎢ ⎥
⎢ 0 0 0 0 0 − Kc Kc + Kdriven − Kdriven ⎥
⎢ ⎥
⎣ 0 0 0 0 0 0 − Kdriven Kdriven ⎦

⎡ ⎤
⎢ ⎥
⎢ ⎥
⎢ C0,1 − C0,1 0 0 0 0 0 0 ⎥
⎢ ⎥
⎢ IE (θ1 )ω + C0,1 ⎥

⎢− C − C1,2 0 0 0 0 0 ⎥
⎢ 0,1 ⎥

⎢ +C1,2 + C1 ⎥

⎢ ⎥
IE (θ2 )ω + C1,2

⎢ ⎥
⎢ 0 − C1,2 − C2,3 0 0 0 0 ⎥
⎢ +C2,3 + C2 ⎥
C =⎢
⎢ 0
⎥, and T = (0, M1, M2, 0, 0, 0, 0,
⎢ 0 − C2,3 C2,3 + C3 − C3 0 0 0 ⎥

⎢ ⎥
⎢ 0 0 0 − C3 C3 + C4 C4 0 0 ⎥
⎢ ⎥
⎢ 0 0 0 0 − C4 C4 + Cc − Cc 0 ⎥
⎢ ⎥
⎢ ⎥
⎢ 0 0 0 0 0 − Cc Cc + Cdriven − Cdriven ⎥
⎢ ⎥
⎢ 0 0 0 0 0 0 − Cdriven Cdriven ⎥
⎢ ⎥
⎣ ⎦

Tload)T.

3. Numerical simulation

Using the Newmark method, Eq. (23) will be numerically simulated to reveal the torsional vibration of the engine crankshaft-
gearbox coupled system varying with the instantaneous angular speed. Take time step dt = 0.0001 s, the sampling frequency is
thus 10 kHz. The simulation procedure is shown in Fig. 7. Start from an initial status, the non-constant inertia of each cylinder IE(θ) and
its derivative to time I’E(θ) are calculated at each time step to form the mass and damping matrices, M and C. The external excitations
from the cylinder and propeller are calculated to form the load vector T. The time varying mesh stiffness of the gear pair Km and the
response-dependent stiffness Kdriven are calculated to form the stiffness matrix K. Then the dynamic responses are evaluated and the

Fig. 8. The second natural frequency of the crankshaft system, the gearbox system and the coupled system.

10
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

Table 5
Maximum and minimum values of the natural frequencies (Hz).
Crankshaft system Gearbox system Coupled system

0 0 0
10.2199/10.2056 12.8456/12.8444 9.2268/9.2242
91.8855/89.5642 481.0334/481.0334 10.8523/10.8390
218.9472/212.5541 816.8027/539.6331 89.3821/87.1720
– – 113.4727/112.7953
– – 219.0022/212.6147
– – 481.0334/481.0334
– – 721.3047/544.7207

whole simulation ends at t > Ttotal.


In frequency domain, the component of f is defined as 1x, and the multiples of f are 2×, 3×, etc.

3.1. Dynamic responses of the coupled system

In this section, the natural frequencies of the healthy coupled system are discussed and the dynamic responses at both the flywheel
and the driven shaft are then presented and compared with each other.

3.1.1. Analysis on natural frequencies


The natural frequencies of the coupled system are evaluated by equation |K-ω2M|=0. Because the coupled system has rigid body
motion, there is zero natural frequency, which is noted as the first order one.
Divide the coupled system to the engine crankshaft and gearbox systems, namely the lumped mass with subscripts 0 to 3 in Fig. 2
are the crankshaft system, the stiffness matrix of which is composed by the top left 4 × 4 elements in K (K3 = 0). The pinion, gear,
coupling and driven shaft constitute the gearbox system and its matrix is composed by the bottom right 4 × 4 elements in K (K3 = 0).
Because of the time varying mass matrix and/or stiffness matrix, the procedure of natural frequency evaluation is similar to Fig. 7, only
that the dynamic equation calculation in each time step is replaced with |K − ω2M| = 0.
Time waveform at the second natural frequency of the engine crankshaft system is presented as the blue curve in Fig. 8. Because of
the term cos2θ in the non-constant inertia IE(θ) = Ie(1 − εcos2θ), IE(θ) will vary with time in a harmonic way and the frequency is twice
of f, whereas the stiffness matrix of the crankshaft system remains constant. So the natural frequency also fluctuates as a harmonic
wave with time.
In Fig. 8, the red curve is the second natural frequency of the gearbox system. Though the mass matrix of the gearbox system is
constant, its stiffness matrix varies with time due to that of the mesh stiffness. As a result, its natural frequency also fluctuates in a same
pattern as Fig. 5.
The second natural frequency of the coupled system is shown in black in Fig. 8. Obviously it is a combination of the harmonic and
saw teeth patterns, contributed by the crankshaft and the gearbox systems respectively.
The maximum and minimum values of the natural frequencies of the crankshaft, gearbox and coupled systems are shown in Table 5.
One finds that the crankshaft and the gearbox system in series decreases the entire stiffness, and then decreases all the natural
frequencies.

(a) (b)
Fig. 9. (a) Time waveforms and (b) spectra of the instantaneous angular speed at the flywheel and driven shaft (f = 20 Hz).

11
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

(a) (b)
Fig. 10. Spectra of the instantaneous angular speed at (a) the flywheel and (b) driven shaft with different rotating frequencies.

3.1.2. Analysis of the torsional vibrations


The torsional vibrations of the crankshaft-gearbox coupled system based on IAS with rotating frequency f = 20 Hz are shown in
Fig. 9, where the blue curve represents the IAS at the flywheel and the red curve at the driven shaft.
In general, IAS at the flywheel is larger in amplitude than that at the driven shaft, as shown in Fig. 9(a). Despite that the external
torques of two cylinders increases when transmitted from the pinion to the gear, IAS at the driven shaft is smaller because the general
stiffness of the gearbox system is larger than that of the crankshaft system.
As shown in Fig. 9(b), frequency components at fdriven and its multiples are prominent in the spectrum of IAS at the flywheel,
amplitudes of which are significantly lower than those at the driven shaft; frequency components at f and its multiples are also evident
in the spectrum of the IAS at the driven shaft, amplitudes of which are significantly lower than those at the flywheel as well. The reason
is that the elastic coupling absorbs the torsional vibration energy, so the vibrations are suppressed when transmitted from the flywheel
to the driven shaft at different frequencies, and vice versa.
Based on the frequency of IAS, dynamic responses of the coupled system can be divided into two groups: the main ones (MA
hereinafter) are the components at the crankshaft rotating frequency f and its multiples with higher amplitudes, which originate from
the external torque of the crankshaft. The minor ones (MI hereinafter) are the components at driven shaft rotating frequency fdriven and
its multiples with lower amplitudes, which come from the load of propeller.
The spectra of IAS at the flywheel and driven shaft in stable states, when f is 20, 24, 28, 32, 36 and 40 Hz, are listed in vertical
direction as shown in Fig. 10, where the colour represent amplitude level. It shows that at the flywheel and driven shaft, the 0.5×

(a) (b)
Fig. 11. (a) Time waveforms and (b) spectra of the instantaneous angular speed at the flywheel and driven shaft (f = 20 Hz, misfire).

12
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

(a) (b)
Fig. 12. Spectra of the instantaneous angular speed at (a) flywheel and (b) driven shaft with different rotating frequencies (Misfire).

components are the largest but decrease with the increasing rotating frequency. The reason is the combustion gas torque Mp, with
frequency is half of f, plays the leading role among the external torque that applied on the crankshaft. When f = 20 Hz, 0.5× is between
the second and the third natural frequencies of the coupled system, the torsional vibration is close to the resonance. As the 0.5× leaves
the two natural frequencies with increasing f, the corresponding amplitude is decreased. The increasing 1.5×, 2× and 2.5× come from
the torque due to the non-constant inertia Mnc and the reciprocating inertia torque Mre, which will increase with rotating frequency
based on Eq. (11) and Eq. (12).
In Fig. 10(b), 0.5×, 1.5×, 2× and 2.5× share the same discipline with those at the flywheel, and the MI increases with f. On the
other hand, MA at the flywheel is larger than MI in amplitude, which makes MI not that obvious in Fig. 10(a). At the driven shaft, MA is
lower, whilst MI is higher, than that at the flywheel, so the external excitations from both the crankshaft and the load are better
reflected, which means IAS monitoring at the driven shaft will not only avoids the intrusive modification of the engine body, but also
exhibits the torsional vibrations of the coupled system in a more obvious way.

3.2. Dynamic responses of the coupled system with misfire

The natural frequencies of the coupled system are not influenced by misfire. The torsional vibration of the coupled system with

Fig. 13. The 2nd natural frequency of coupled system with breathing slant crack on the driven shaft.

13
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

(a) (b)

Fig. 14. (a) Time waveforms and (b) spectra of the instantaneous angular speed at the flywheel and driven shaft (f = 20 Hz, slant crack).

100% misfire of the first cylinder is investigated in this section.


Time waveforms and spectra of IAS at the flywheel and driven shaft are shown in Fig. 11. The amplitudes at two places decrease a
lot [49,50], and the fluctuation at the top of the waveform at the flywheel also decreases compared to that in Fig. 9(a). We also find that
the curve of the driven shaft is smoother, which reflects the crankshaft structural characteristics.
The crankshaft rotates with a period of 4π rad. During every period, it goes through air-in, compression, combustion and air-out,
each of which is in correspondence to rotation angle of π rad of the crank. On the other hand, the phase difference of π rad between two
cranks enlarges the difference between the excitations that applied to cranks, and then increases the torsional vibrations. Due to the
combustion gas torque Mp is zero in misfire case, the torsional vibrations are lower.
In Fig. 12(a), spectra of IAS at the flywheel show similar discipline with those in Fig. 10(a). The load from the propeller is not
affected by misfire, thus MI are same as those in Fig. 10. With decreasing MA, MI in Fig. 12(a) is more obvious.

3.3. Dynamic responses of the coupled system with slant crack

Equations in Section 2.4 shows that the stiffness of the driven shaft is dependent on the torsional vibrations, that is the stiffness of
the driven shaft and the dynamic responses of the coupled system can influence with each other. Dynamic response of the coupled
system with slant crack is analysed in this section, regardless of the misfire in the cylinders.

3.3.1. Analysis of natural frequencies


The second natural frequency of the coupled system with breathing slant crack on the driven shaft when f = 20 Hz is shown in
Fig. 13.
Compared with the black curve in Fig. 8, when the slant crack on the driven shaft opens, the stiffness will suddenly decrease, which
also results in reduction of the total stiffness and natural frequencies of the coupled system. The time varying stiffness matrix K is not
same as that in Section 3.1.1, therefore the curve in Fig. 13 shows no regularity.

3.3.2. Analysis of the torsional vibrations


Comparison between Fig. 14 and Fig. 9 indicates that the breathing slant crack has little effect on the IAS at the flywheel when f =
20 Hz.
In Fig. 14(a), the peaks of IAS at the driven shaft are over 150◦ /s, some troughs are lower than − 150◦ /s. Fluctuations in small
amplitudes appear in the time waveforms, especially near the peaks and troughs. This phenomenon means when the amplitude of
torsional vibration is larger, the sign of φdriven− φc changes rapidly, resulting in frequent crack opening and closing. At same time, MI at
the driven shaft increases significantly compared to those in Fig. 9(b), see Fig. 14(b).
Similar to the case f = 20 Hz, IAS at the flywheel with breathing slant crack on the driven shaft and different rotating frequency
show no difference with healthy ones. MI at the driven shaft are more prominent and the amplitudes increase significantly [51]. At f =
36 Hz, the amplitude of 6fdriven is higher than 300◦ /s. The reason is that the sign of φdriven− φc changes more frequently, so are the crack
opening and closing at the peaks and troughs. As a result, the amplitudes of the high frequency components increase. Fig. 15 also shows
that the difference between the dynamic response of coupled system under crack and healthy conditions is more obvious at the driven
shaft. So the amplitude changes of frequency components related to fdriven can be used to diagnose the crack fault at the driven shaft. At
f = 20 Hz, amplitude of 5fdriven at the driven shaft increases from 1.85◦ /s to 13.04◦ /s with increases of depth of crack from 0 to 3 mm, as
shown in Table. 6.

14
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

(a) (b)

Fig. 15. Spectra of the instantaneous angular speed at (a) the flywheel and (b) the driven shaft with different rotating frequencies (Slant crack).

Table 6
Amplitude of 5fdriven at the driven shaft under different crack depths (f =
20 Hz).
Depth of crack (mm) Amplitude of 5fdriven (◦ /s)

0 1.85
0.11 2.12
0.3 3.54
0.5 5.89
1 8.33
2 11.27
3 13.04

Table 7
Amplitude of IAS under different conditions (◦ /s).
Condition Flywheel Driven Shaft

0.5× 5fdriven 0.5× 5fdriven

Normal 20 Hz 182.30 0.12 110.70 1.85


40 Hz 122.5 0.92 39.58 3.15
Misfire 20 Hz 134.00 0.12 81.32 1.85
40 Hz 86.94 0.92 28.17 3.15
Crack 20 Hz 182.20 0.15 111.70 13.04
40 Hz 122.5 1.86 40.21 221.1

3.4. Summary

Numerical results of the torsional vibrations of the engine crankshaft-gearbox coupled system under normal, engine misfire and
breathing slant crack conditions reveal three features of IAS at the driven shaft.

1) For normal condition, amplitude of MA at the driven shaft is relatively lower and that of MI higher compared to those at the
flywheel, which makes them easier to be observed.
2) For misfire condition, amplitude of MA decreases and MI remains constant. Change of the MA amplitude can be adopted as the
criteria of misfire.
3) For crack condition, amplitude of the time waveform of IAS will fluctuate slightly near the peaks and troughs, and the amplitude of
MI in the spectrum increase significantly.

Furthermore, based on IAS we can also obtain the basic law of condition monitoring of the coupled system at the driven shaft.
Amplitudes of 0.5× and 5fdriven in the spectra of IAS at the flywheel and driven shaft under three conditions (namely Fig. 9, Fig. 11 and
Fig. 14) when f = 20 and 40 Hz are listed in Table. 7. Take f = 20 Hz as an example, for IAS at the flywheel, only 0.5× decreases from
around 182◦ /s to 134.00◦ /s when there exists misfire in the cylinder, while 5fdriven changes from 0.12◦ /s to 0.15◦ /s in three conditions.

15
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

Fig. 16. Flow chart from measurement to fault diagnosis.

Fig. 17. Sensors at (a) the flywheel, (b) driven shaft and (c) encoder.

For IAS at the driven shaft, the amplitude of 5fdriven increases from 1.85◦ /s to 13.04◦ /s when there exists a crack, besides the change of
0.5× (from 110.70◦ /s to 81.32◦ /s) in misfire condition.
To carry out condition monitoring, the basic IAS data of the driven shaft under normal condition should be experimentally obtained
at different rotating frequencies. Note that due to some unstable parameter variations, the IAS at a rotating frequency can also vary
within a reasonable range. When MA decrease below the minimum value of the normal range, there exists misfire in the cylinder. When
MI increase above the maximum value of the normal range and the amplitude of time waveform fluctuates slightly near the peaks and
troughs, there exist breathing slant crack at the driven shaft. Flow chart from measurement to fault diagnosis is shown in Fig. 16.

4. Experiment validation

4.1. Introduction to the test rig

Experiment procedure is to run the test rig, shown in Fig. 1, for ten minutes to warm up at idling speed, then increase the rotating
frequency to the set value and hold for 5 min. Response of the last 40 s are recorded for analysis. After obtaining the maximum rotation

16
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

Fig. 18. Time waveform of the voltage output by the photoelectric sensor.

(a) (b)

Fig. 19. (a) Time waveform and (b) spectrum of the instantaneous angular speed at the flywheel (f = 18.15 Hz).

frequency data, stop the engine and cool down for 2 h. Repeat this process three times and the middle results are shown in the following
section.
An electromagnetic transducer is installed at the flywheel by drilling a hole on the engine body, shown inside the blue box in Fig. 17
(a). A photoelectric sensor, inside the green dashed box in Fig. 17(b), is used to collect the rotation signal of a 3D printed encoder. The
encoder, inside the red dashed box in Fig. 17(b), has 60 holes along its edge as shown in Fig. 17(c). There are 86 teeth on the side of the
flywheel. When the gap between the flywheel teeth and the sensor increases, the output voltage of the electromagnetic sensor will
change. In this way, the sensor converts the rotation of flywheel into a voltage signal, which is similar to the photoelectric sensor.
Fig. 18 shows the time waveform of the voltage output by the photoelectric sensor. A period starts when the voltage exceeds the
threshold (the photoelectric sensor is 3.5 V) and ends when the voltage is higher than the threshold again. During this period, the
encoder rotates an angle of 2π/60. Dividing 2π/60 by the length of the time period results in the rotating speed of the encoder. The
sampling frequency here is 102.4 kHz in coordination with other sensors.
The measurement noise in the experiment comes from the fluctuation of the rotating frequency of the engine crankshaft, the
integrity of the encoder and data processing method. The integrity of flywheel and encoder has been thoroughly checked before and
after the experiment, and no fault was found. As for the data processing method, the highest order of harmonics, Hana, should satisfy
Hana < Nteeth/holes/4, where Nteeth/holes is the number of teeth/holes on the flywheel or the encoder [52]. We took Hana = 15 and the key

17
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

(a) (b)
Fig. 20. (a) Time waveform and (b) spectrum of the instantaneous angular speed at the driven shaft (f = 18.15 Hz).

(a) (b)
Fig. 21. Spectra of the instantaneous angular speed at (a) the flywheel and (b) the driven shaft (f = 22.35 Hz).

frequency components to detect misfire and crack are 0.5× and 5fdriven respectively, so the transformation and the leakage have little
influence on the results. In summary, the measurement noise that may influence the results is mainly the fluctuation of the rotating
frequency of the engine crankshaft.

4.2. Experimental validation

The time waveforms of IAS at the flywheel and the driven shaft for f = 18.15 Hz are shown in Fig. 19(a) and Fig. 20(a). The last half
second data are enlarged in red for displaying clearly the response wave of driven shaft. Spectra shown in Fig. 19(b) and Fig. 20(b)
show that the main components of IAS at flywheel are 0.5×, 1.5× and 3×.
Spectra of IAS at the flywheel and driven shaft, at the rotating frequency of crankshaft of 22.35, 26.75, 33.98 and 38.32 Hz
respectively, are shown in Figs. 21–24. In good accordance with the simulation results, one finds that the amplitude of the flywheel
corresponding to 0.5× is always the highest, which decreases from 188.35 to 112.88◦ /s with increase of the rotating frequency.
Though the MI can be observed at the flywheel in simulation in Section 3.1.2, it cannot be observed in the experiment due to damping
effect of lubricating oil inside the gearbox.
In the spectra of IAS of the driven shaft, MA are obvious and 0.5× decreases while MI increase with increasing rotating frequency,
which also verifies the simulation results. It should be noted that in simulation, the components with twice, three and four times the
frequency fdriven are clear, whereas in experiment the prominent components become six, eight and ten times the fdriven. The reason is
the propeller in experiment rotates in a tank with circulating water, rather than in the open area represented by Eq. (22).

18
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

(a) (b)
Fig. 22. Spectra of the instantaneous angular speed at (a) the flywheel and (b) the driven shaft (f = 26.75 Hz).

(a) (b)
Fig. 23. Spectra of the instantaneous angular speed at (a) the flywheel and (b) the driven shaft (f = 33.98 Hz).

Besides, the amplitude of 5fdriven at the driven shaft is 2.12◦ /s, when f = 18.15 Hz, in the experiment. Such value is 1.85◦ /s for α = 3
mm and 2.12◦ /s for α = 0.11 mm in the simulation, as listed in Table 6. In other words, the minimum size of the crack that can be
identified by the proposed technique is about 0.11 mm.
Based on the experiment results, advantages for installing the encoder at the driven shaft to monitor IAS are,

1) There is no need to drill holes in the engine body.


2) The encoder is easy to install because there is enough space, and away from the high temperature and vibrating engine during
operation.
3) The IAS on the driven shaft includes the information from the crankshaft and propeller load. According to this information, the
crankshaft and gearbox can be monitored at the same time, while the IAS on the flywheel cannot reflect the influence of external
load on the torsional vibration of the engine crankshaft gearbox coupling system.

5. Conclusions

Based on simulation, analysis and comparisons of the instantaneous angular speed at the flywheel and driven shaft, the torsional
vibration of the diesel engine crankshaft-gearbox coupled system under three conditions, normal, misfire and breathing slant crack on
the driven shaft is numerical investigated. The results of normal condition are also verified by experiment. Conclusions are summarized

19
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

(a) (b)
Fig. 24. Spectra of the instantaneous angular speed at (a) the flywheel and (b) the driven shaft (f = 38.32 Hz).

as follows.

1) For the crankshaft, the non-constant inertia varies in harmonic way and the stiffness is constant, so its natural frequencies fluctuate
within a small range in a harmonic way. For the gearbox, the mesh stiffness is time varying and the inertia is constant, so its natural
frequencies fluctuate in the same way of stiffness. Natural frequencies of the coupled system are the combination of these two
subsystems.
2) Under normal condition, the main part of torsional vibration of the coupled system comes from the driving torque of the crankshaft,
and its frequency is a multiple of the crankshaft rotation frequency, and the secondary part comes from the load of the propeller.
The IAS at the flywheel has a major part and a minor part, and the latter has a low amplitude. In the IAS of the driven shaft, these
two components are the same magnitude, so it is a better place to monitor the condition of the crankshaft and gearbox at the same
time.
3) The amplitude of the main part at the driven shaft decreases significantly due to the engine structural characteristics, and it is easier
to observe two parts.
4) Small fluctuations appear near the peaks and troughs of the IAS at the driven shaft in time domain, and the minor part increases
significantly.
5) The experiment indicates that the minor part of the IAS at the flywheel is negligible due to the damping effect of lubricating oil
inside the gearbox. But the two parts are both prominent at the driven shaft and reflect similar regularity when rotating frequency
changes, which means the IAS sensor at the driven shaft can monitor the crankshaft and the gearbox at the same time.

Using the IAS dynamic characteristics obtained at the driven shaft, the misfire and crack status of the engine crankshaft gearbox can
be monitored according to the following procedures.

1) Collect the basic IAS data at the driven shaft under normal condition with different rotating frequencies.
2) Apply Fast Fourier Transform to the collected IAS signals. When the amplitudes of the components at multiples of rotating fre­
quency of the crankshaft drops to a lower value than normal case, it can be concluded that there is misfire in an engine cylinder.
3) When amplitudes of the components at multiples of the rotating frequency of the driven shaft exceeds a higher value compared to
normal condition, and fluctuations in small amplitude appear near the peaks and troughs in the time waveform of the IAS signals,
there exists breathing slant crack at the driven shaft.

Declaration of Competing Interest

The authors declare the following financial interests/personal relationships which may be considered as potential competing in­
terests: Qian Ding reports financial support was provided by National Natural Science Foundation of China. Qian Ding reports financial
support was provided by Natural Science Foundation of Tianjin City. Kun Wu reports financial support was provided by China
Scholarship Council.

Acknowledgements

This work was supported by the Natural Science Foundation of Tianjin City through the grants (19JCZDJC38800).

20
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

References

[1] P. Charles, J.K. Sinha, F. Gu, L. Lidstone, A.D. Ball, Detecting the crankshaft torsional vibration of diesel engines for combustion related diagnosis, J. Sound Vib.
321 (3-5) (2009) 1171–1185.
[2] W. Homik, Diagnostics, maintenance and regeneration of torsional vibration dampers for crankshafts of ship diesel engines, Polish Mar. Res. 17 (2010) 62–68.
[3] W. Homik, Damping of torsional vibrations of ship engine crankshafts – general selection methods of viscous vibration damper, Polish Mar. Res. 18 (2011)
43–47.
[4] V.N. Nikishin, A.P. Pavlenko, K.N. Svetlichnyi, V.S. Gol’makov, Analysis of torsional crankshaft oscillations in a diesel engine on the basis of cylinder-block
vibration, Russ. Eng. Res. 33 (2014) 687–691.
[5] R. Potenza, J.F. Dunne, S. Vulli, D. Richardson, A model for simulating the instantaneous crank kinematics and total mechanical losses in a multicylinder in-line
engine, Int. J. Engine Res. 8 (4) (2007) 379–397.
[6] B. Chiliński, M. Zawisza, Analysis of bending and angular vibration of the crankshaft with a torsional vibrations damper, J. Vibroeng. 18 (2016) 5353–5363.
[7] H. Karabulut, Dynamic model of a two-cylinder four-stroke internal combustion engine and vibration treatment, Int. J. Engine Res. 13 (2012) 616–627.
[8] S. Ni, Y. Guo, W. Li, D. Wang, Z. Shuai, D. Yu, Effect of advanced injection angle on diesel engine shaft torsional vibration, Int. J. Engine Res. 22 (2020)
1457–1464.
[9] Z. Liu, K. Wu, Q. Ding, J.X. Gu, Engine misfire diagnosis based on the torsional vibration of the flexible coupling in a diesel generator set: simulation and
experiment, J. Vib. Eng. Technol. 8 (1) (2020) 163–178.
[10] Q. Zhang, J. Duan, S. Zhang, Y. Fu, Nonlinear dynamic modeling for a diesel engine propeller shafting used in large marines, Chin. J. Mech. Eng. 27 (2014)
937–948.
[11] M. Pfabe, C. Woernle, Reducing torsional vibrations by means of a kinematically driven flywheel — Theory and experiment, Mech. Mach. Theory 102 (2016)
217–228.
[12] L. Sun, F. Luo, T. Shang, H. Chen, A. Moro, Research on torsional vibration reduction of crankshaft in off-road diesel engine by simulation and experiment,
J. Vibroeng. 20 (1) (2018) 345–357.
[13] W. Li, Y. Guo, X. Lu, X. Ma, T. He, D. Zou, Tribological effect of piston ring pack on the crankshaft torsional vibration of diesel engine, Int. J. Engine Res. 16 (7)
(2015) 908–921.
[14] Q. Wang, P. Hu, Y. Zhang, Y. Wang, X. Pang, C. Tong, A Model to Determine Mesh Characteristics in a Gear Pair with Tooth Profile Error, Adv. Mech. Eng. 6
(2015) 1–10.
[15] 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 (4) (2008) 691–705.
[16] W. Li, D. Pang, Analytical investigation on geared rotor system with multi-body fault based on finite element method, J. Vib. Control 25 (2018) 408–422.
[17] F. Chaari, T. Fakhfakh, M. Haddar, Analytical modelling of spur gear tooth crack and influence on gearmesh stiffness, Eur. J. Mech. A. Solids 28 (2009) 461–468.
[18] P. Sainsot, P. Velex, O. Duverger, Contribution of Gear Body to Tooth Deflections—A New Bidimensional Analytical Formula, J. Mech. Des. 126 (2004)
748–752.
[19] 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.
[20] 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.
[21] Y. Luo, N. Baddour, M. Liang, Effects of gear center distance variation on time varying mesh stiffness of a spur gear pair, Eng. Fail. Anal. 75 (2017) 37–53.
[22] Z.G. Chen, Y.M. Shao, T.C. Lim, Non-linear dynamic simulation of gear response under the idling condition, Int. J. Automot. Technol. 13 (2012) 541–552.
[23] Y. Luo, N. Baddour, G. Han, F. Jiang, M. Liang, Evaluation of the time-varying mesh stiffness for gears with tooth spalls with curved-bottom features, Eng. Fail.
Anal. 92 (2018) 430–442.
[24] N.L. Pedersen, M.F. Jørgensen, On gear tooth stiffness evaluation, Comput. Struct. 135 (2014) 109–117.
[25] W. Yu, Y. Shao, C.K. Mechefske, The effects of spur gear tooth spatial crack propagation on gear mesh stiffness, Eng. Fail. Anal. 54 (2015) 103–119.
[26] H. Ma, R. Song, X. Pang, B. Wen, Time-varying mesh stiffness calculation of cracked spur gears, Eng. Fail. Anal. 44 (2014) 179–194.
[27] B. Liu, C. Zhao, F. Zhang, T. Cui, J. Su, Misfire detection of a turbocharged diesel engine by using artificial neural networks, Appl. Therm. Eng. 55 (2013) 26–32.
[28] X. Xu, Z. Liu, J. Wu, J. Xing, X. Wang, Misfire Fault Diagnosis of Range Extender Based on Harmonic Analysis, Int. J. Automot. Technol. 20 (2019) 99–108.
[29] S.B. Devasenapati, V. Sugumaran, K.I. Ramachandran, Misfire identification in a four-stroke four-cylinder petrol engine using decision tree, Expert Syst. Appl.
37 (2010) 2150–2160.
[30] S.K. Tummala, Ł. Grabowski, P. Karpiński, K. Pietrykowski, P.B. Bobba, Simulation of an aircraft radial engine misfire detection, E3S Web Conf. 87 (2019) 1–6.
[31] C. Xu, S. Li, F. Cao, X. Qiu, Misfire Detection Based on Generalized Force Identification at the Engine Centre of Gravity, IEEE Access 7 (2019) 165039–165047.
[32] D. Jung, E. Frisk, M. Krysander, A flywheel error compensation algorithm for engine misfire detection, Control Eng. Pract. 47 (2016) 37–47.
[33] S. Singh, S. Potala, A.R. Mohanty, An improved method of detecting engine misfire by sound quality metrics of radiated sound, Proc. Inst. Mech. Eng. Part D: J.
Automob. Eng. 233 (12) (2019) 3112–3124.
[34] J. Chen, R. Bond Randall, Improved automated diagnosis of misfire in internal combustion engines based on simulation models, Mech. Syst. Sig. Process. 64–65
(2015) 58–83.
[35] N. Cavina, G. Cipolla, F. Marcigliano, D. Moro, L. Poggio, A methodology for increasing the signal to noise ratio for the misfire detection at high speed in a high
performance engine, Control Eng. Pract. 14 (3) (2006) 243–250.
[36] S. Helm, M. Kozek, S. Jakubek, Combustion Torque Estimation and Misfire Detection for Calibration of Combustion Engines by Parametric Kalman Filtering,
IEEE Trans. Ind. Electron. 59 (2012) 4326–4337.
[37] A. Sharma, V. Sugumaran, S. Babu Devasenapati, Misfire detection in an IC engine using vibration signal and decision tree algorithms, Measurement 50 (2014)
370–380.
[38] L. Rubio, J. Fernández-Sáez, A new efficient procedure to solve the nonlinear dynamics of a cracked rotor, Nonlinear Dyn. 70 (2012) 1731–1745.
[39] L. Xiang, Y. Zhang, A. Hu, Crack characteristic analysis of multi-fault rotor system based on whirl orbits, Nonlinear Dyn. 95 (2018) 2675–2690.
[40] Q. Han, F. Chu, Dynamic instability and steady-state response of an elliptical cracked shaft, Arch. Appl. Mech. 82 (2011) 709–722.
[41] Q. Han, F. Chu, Parametric instability of a Jeffcott rotor with rotationally asymmetric inertia and transverse crack, Nonlinear Dyn. 73 (2013) 827–842.
[42] R. Ramezanpour, M. Ghayour, S. Ziaei-Rad, A novel method for slant crack detection in rotors based on turning in two directions, Arch. Appl. Mech. 83 (2012)
783–798.
[43] X. Lei, G. Zhang, J. Chen, S. Xigeng, G. Dong, Simulation on the motion of crankshaft with a slant crack in crankpin, Mech. Syst. Sig. Process. 21 (2007)
502–513.
[44] A.K. Darpe, K. Gupta, A. Chawla, Coupled bending, longitudinal and torsional vibrations of a cracked rotor, J. Sound Vib. 269 (2004) 33–60.
[45] B. Zhang, Y. Li, Six degrees of freedom coupled dynamic response of rotor with a transverse breathing crack, Nonlinear Dyn. 78 (2014) 1843–1861.
[46] Y. Huang, S. Yang, F. Zhang, C. Zhao, Q. Ling, H. Wang, Non-linear torsional vibration characteristics of an internal combustion engine crankshaft assembly,
Chinese Journal of, Mech. Eng. 25 (2012) 797–808.
[47] T. Zhang, Z. Chen, W. Zhai, K. Wang, Establishment and validation of a locomotive–track coupled spatial dynamics model considering dynamic effect of gear
transmissions, Mech. Syst. Sig. Process. 119 (2019) 328–345.
[48] A.J. Sørensen, Ø.N. Smogeli, Torque and power control of electrically driven marine propellers, Control Eng. Pract. 17 (2009) 1053–1064.
[49] Q. Song, W. Gao, P. Zhang, J. Liu, Z. Wei, Detection of engine misfire using characteristic harmonics of angular acceleration, Proc. Inst. Mech. Eng. Part D: J.
Automob. Eng. 233 (14) (2019) 3816–3823.

21
K. Wu et al. Mechanical Systems and Signal Processing 173 (2022) 109052

[50] A. Hmida, A. Hammami, F. Chaari, M. Ben Amar, M. Haddar, Effects of misfire on the dynamic behavior of gasoline Engine Crankshafts, Eng. Fail. Anal. 121
(2021) 1–19.
[51] C. Liu, D. Jiang, Torsional vibration characteristics and experimental study of cracked rotor system with torsional oscillation, Eng. Fail. Anal. 116 (2020) 1–16.
[52] S.D. Yu, X. Zhang, A data processing method for determining instantaneous angular speed and acceleration of crankshaft in an aircraft engine–propeller system
using a magnetic encoder, Mech. Syst. Sig. Process. 24 (2010) 1032–1048.

22

You might also like