Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Mechanism and Machine Theory 99 (2016) 19–36

Contents lists available at ScienceDirect

Mechanism and Machine Theory


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

Investigation on the influence of the cavitation boundaries on


the dynamic behavior of planar mechanical systems with
hydrodynamic bearings
Gregory Bregion Daniel ⁎, Tiago Henrique Machado, Katia Lucchesi Cavalca
Laboratory of Rotating Machinery, Faculty of Mechanic Engineering, UNICAMP, 200, Rua Mendeleyev, PO Box 6122, Campinas, CEP 13083-970, SP, Brazil

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

Article history: The characterization of lubricated revolute joints configures a topic of great interest in
Received 21 August 2015 mechanisms and machines since there is a significant demand for its proper design in many
Received in revised form 5 November 2015 applications. Therefore, this work proposes a solution for the evaluation of the hydrodynamic
Accepted 9 November 2015
force components of the journal bearings obtained from the integration of the Reynolds
Available online 18 January 2016
equation for infinitely short bearings. The main difficulty in obtaining these hydrodynamic
forces lies on adequately defining the boundary conditions of the Reynolds equation. Differently
Keywords: from the boundary conditions proposed in classical literature, this work presents an approach
Hydrodynamic lubrication
to determine the integration boundaries, i.e., the angular position in the bearing reference system
Hydrodynamic forces
where the pressure is null. Therefore, the hydrodynamic forces are calculated taking into account
Short bearings
Slider–crank mechanisms only the positive pressure. These angles are time dependent, and its evaluation is very important
for a more accurate estimation of the hydrodynamic force components. The computer simulations
are accomplished for different length–diameter ratios (L/D). From the global results, it can be
concluded that the proposed approach is promising to efficiently represent the hydrodynamic
conditions in bearings with L/D ratios up to 0.5.
© 2016 Elsevier Ltd. All rights reserved.

1. Introduction

A topic of great interest in mechanisms and machines is the characterization of joints, which directly affects the dynamic
behavior of the mechanical system. Generally, these joints are designed to operate with some lubrication since the film formed
by the lubricant reduces friction and wear provides load-carrying capacity, adds damping to dissipate undesirable mechanical
vibration, and increases the heat transfer to reduce the temperature [1]. Therefore, the proper description of lubricated revolute
joints, the well-known journal bearings, in multibody mechanical systems is required to achieve better models and, hence, an
improved understanding of the dynamic performance of machines. This study configures a topic of significant impact due to
the demand for the proper design of the journal bearings in many industrial applications [2].
The lubrication model applied to journal bearings must be carefully investigated when the clearance is considered. In the
hydrodynamic bearings, the clearance is filled with a lubricant that generates a pressure distribution and, consequently, hydrody-
namic forces actuating in the components of the mechanical system. These hydrodynamic forces include both squeeze and wedge
effects that oppose to the journal motion, and they are obtained by integrating the pressure distribution evaluated through the
Reynolds equation written for the dynamic regime [1]. The hydrodynamic forces are nonlinear functions of the journal center
position and of its velocity with reference to the bearing center. In the dynamic regime of a journal bearing, the journal center

⁎ Corresponding author. Tel.: +55 19 3521 3178.


E-mail address: gbdaniel@fem.unicamp.br (G.B. Daniel).

http://dx.doi.org/10.1016/j.mechmachtheory.2015.11.019
0094-114X/© 2016 Elsevier Ltd. All rights reserved.
20 G.B. Daniel et al. / Mechanism and Machine Theory 99 (2016) 19–36

has an orbit situated within a circle radius, equal to the radial clearance. In a simple way, the hydrodynamic forces built up by the
lubricant fluid are evaluated from the knowledge of the system variables and then included into the equations of motion of the
multibody system [2].
Over the years, several authors have been investigating, theoretically and experimentally, the dynamic influence of lubrication
at the clearance of journal bearings in the framework of multibody systems. Initially, they attempted to solve the Reynolds
equation by means of analytical and numerical methods. In the late 40s, more advanced calculation methods were developed,
and in the 50s, with the advent of the computational mechanics, the numerical solutions of Reynolds equation become feasible,
leading to perform more complete models.
However, the exact solution of the Reynolds equation is difficult to obtain and, in general, requires a considerable numerical
effort [3], especially when the journal bearings are part of a complex mechanical system. Nevertheless, it is possible to solve
the equation analytically by setting to zero either the term that takes into account the oil leakage at the ends or the term of
the circumferential flow of the bearing. These solutions correspond to those for infinitely short and infinitely long journal
bearings, respectively.
The most current models used to describe the effect of bearings hydrodynamic lubrication in multibody systems were
proposed by [1,3]. Pinkus and Sternlicht [1] developed models for infinitely long journal bearings considering different boundaries
and dynamic conditions. In Frêne et al. [3], formulations for both long and short bearings were described considering two different
boundaries conditions, namely, the Sommerfeld's and the Gümbel's conditions.
The models established in [1,3] have been widely applied in various works, which have included the lubrication action at
the joints clearance in computer simulations of multibody systems. Flores et al. [4] accomplished a dynamic analysis of a
mechanical system with lubricated journal bearings, applying a hydrodynamic lubrication model to an infinitely long journal
bearing, taking into account the hydrodynamic forces due to only squeeze effects. However, for high angular velocities, the
simple squeeze approach was not valid and, consequently, other approach should be applied. In the same year, Flores
et al. [5] applied the infinitely long journal bearing model to a slider–crank mechanism, in which a lubricated revolute
joint connects the connecting rod and slider, to discuss the assumptions and procedures adopted. The results for the hydro-
dynamic lubrication model were compared with those obtained with ideal joints, showing a quite good match, meaning that
the use of lubricant at the machine joints is an effective way to ensure better performance. In the same field, Ashaer et al. [6] used a
general methodology for modeling lubricated forces produced by the lubricant in a dynamically loaded long journal bearing in
multibody mechanical systems. The proposed model predicts that the reaction moment is within the same order of magnitude of
using perfect joint.
Flores et al. [7] continued the study on dynamics of mechanical systems including joints with clearance and a hydrodynamic
lubrication model based on the solution of the Reynolds equation, considering the infinitely long journal bearing condition,
besides squeeze and wedge effects. In order to compare the different models of journal bearings, Flores et al. [8] applied the
models proposed by Frêne [3] in a multibody system with lubricated revolute joints. Thus, the influence of the lubrication
model on the dynamic behavior of the system can be observed. Four models were considered in the computer simulations,
being two models for infinitely short bearing (Gümbel and Sommerfeld boundaries conditions) and two models for infinitely
long bearing (Gümbel and Sommerfeld boundaries conditions). The results obtained in these simulations showed the strong
influence of the bearing model in the dynamic behavior of the mechanical system.
Tian et al. [9] simulated the behavior of planar flexible multibody systems with clearance and lubricated revolute joints. For
lubricated revolute joints, the hydrodynamic forces were calculated as proposed by Flores et al. [8]. However, aiming to consider
possible cavitation effects, some modifications were done in the lubrication model as proposed by Ravn et al. [10].
Among the different approaches used to design dynamically loaded journal bearings, the mobility method has an important
role and is widely applied. Developed by Booker [11,12], the concept of this graphical method is useful to predict the journal
center orbit, in the time domain, using the mobility maps. Flores et al. [13] used the analytical mobility method to analyze journal
bearings subjected to dynamic loads and included it in a general computational program that was developed for the dynamic

Fig. 1. Basic journal bearing geometry. (a) Lateral View; (b) axial view.
G.B. Daniel et al. / Mechanism and Machine Theory 99 (2016) 19–36 21

analysis of general mechanical systems. Choi et al. [14] proposed an analytical approach to tolerance optimization for
planar mechanisms with lubricated joints based on mechanical error analysis. The mobility method was applied to consider the
lubrication effects at joints. In this case, planar mechanisms were stochastically defined by using the clearance vector model for
mechanical error analysis.
Daniel and Cavalca [15] analyzed the dynamic of a slider–crank mechanism with hydrodynamic lubrication in the connecting
rod-slider joint clearance. The hydrodynamic lubrication model was previously developed by Bannwart et al. [16], considering the
infinitely long journal bearing condition. Moreover, the pressure distribution was obtained due to wedge effects and considering
the acceleration of the bearing due to the linear motion of the slider.
In this context, Machado et al. [2] studied the effect of the lubricated revolute joint parameters and hydrodynamic force
models on the dynamic response of planar multibody systems. Three different hydrodynamic force models were investigated,
being the Pinkus and Sternlicht model for infinitely long journal bearings [1] and the Frêne et al. models for infinitely long and
short journal bearings [3]. According to the results obtained in the simulations, the hydrodynamic force models significantly
changed the dynamic characteristics of the multibody systems.
Recently, some studies have applied numerical approaches to calculate the acting forces of lubricated revolute joints in
multibody system. Wang et al. [17] analyzed the revolute joint as a contact problem from interference fitted bearing and
pre-loaded bolts, in order to study the bearings deformation in the connecting rod. The computer simulations of this work
were performed by finite element method using the ABAQUS' solver. In following, Wang et al. [18] calculated the pressure
distribution and the hydrodynamic forces for solving Reynolds equation by finite difference method. The numerical analyses
accomplished in this work showed the lubrication conditions in the connecting rod bearing, considering the effects of deformation
(elastic and rigid bearing) and shape (circular and non-circular bearing). Tian et al. [19] studied the elastohydrodynamic (EHD)
lubrication of cylindrical joints in multibody system. Also in this work, the Reynolds equation was solved in its complete form
by the finite difference method, and the hydrodynamic forces were then used to calculate the connecting rod deformation. The
results obtained in this work were compared with the results obtained by the software ADINA with good agreement. Tian
et al. [20] also studied the coupling dynamics of a geared multibody system supported by EHD lubricated cylindrical joints. The
influence of the lubrication conditions (HD and EHD), the gear mesh force models and the clearance size on the system responses
were also analyzed.
Other studies (Schwab et al. [21] and Erkaya et al. [22]) included the effect of the elasticity of some components of mechanical
systems along with the effects of lubrication in the revolute joints. Finally, Flores and Lankarani [23] modeled and simulated
spatial multibody systems with rigid-lubricated spherical joints. The squeeze film action, due to the relative approaching motion
between the mechanical joint elements, was considered utilizing the lubrication theory associated with the spherical bearings. It
was observed that the system's performance with lubricant effect presents fewer peaks in the kinematic and dynamic outputs,
when compared with those from the dry contact joint model.
On the other hand, the numerical approach is able to evaluate the hydrodynamic forces acting in the internal wall of the
bearing. Moreover, the solution of the complete form of the Reynolds equation leads to more accurate results, although it can
be unfeasibly in some cases due to the high computational costs.
One of the main problems in obtaining the solutions for journal bearing lubrication in dynamic conditions lies not only in
solving the differential equations but also in properly defining the boundary conditions of the Reynolds equation. In dynamically
loaded journal bearings, the evaluation of the force components obtained from the integration of the Reynolds equation only over
the positive pressure regions, assuming the negative pressure as null, involves finding the zero points, that is, the angle for which
a positive pressure begins and the angle for which the pressure is null. For the case of a steady-state journal bearing, half

Fig. 2. Perspective view of a basic journal bearing.


22 G.B. Daniel et al. / Mechanism and Machine Theory 99 (2016) 19–36

Sommerfeld approach, these angles are assumed to be 0 and π, respectively. However, for a dynamically loaded journal bearing,
these angles are time dependent and the evaluation of the force components involves a good deal of mathematical manipulation
[2]. The details in the treatment of these angles are described in the work by Pinkus and Sternlicht for infinitely long journal
bearings [1]. However, there are no studies in the literature dealing with this issue for infinitely short bearing. Within this context,
the present work is inserted.
The main purpose of this work is to propose a solution for the evaluation of the hydrodynamic force components obtained
from the integration of the Reynolds equation for infinitely short bearings. Differently from the boundary conditions proposed
by Sommerfeld and Gümbel [3], in which full film (0 to 2π) and half film (0 to π) are respectively considered in the integration
of the pressure distribution, this work presents an approach to determine the integration boundaries, i.e., the angular position in
the bearing where the pressure is null. Therefore, the hydrodynamic forces are calculated taking into account only the positive
pressure. As previously mentioned, these angles are time dependent, and its evaluation is very important for a more accurate
estimation of the hydrodynamic force components.
Effects of temperature on viscosity as well as the links flexibility, even quite important, were not considered in the present
analysis due to the focus on pressure distribution shifting regarding the local reference, namely, the maximum oil film thickness,
depending on the boundary conditions assumptions.
The analysis of the results from the proposed solution is compared with the classical numerical solution of the
Reynolds equation and two analytical solutions presented by Frêne et al. [3] in order to verify the accuracy and the sensitivity
of the proposed solution. Moreover, the computer simulations were accomplished for different length–diameter ratios (L/D),
aiming to establish the range with better agreement regarding the solution for finite bearing (complete form of
Reynolds equation).

2. Materials and methods

The theory applied in this paper can be basically divided into five parts: (1) the Reynolds equation for a dynamically loaded
journal bearing, (2) the solution of the Reynolds equation for infinitely short bearings and the two models for the hydrodynamic
forces proposed by Frêne et al. [3], (3) the proposed approach to estimate the hydrodynamic forces for infinitely short bearings,
(4) the numerical solution of the classical Reynolds equation by Finite Volume Method (FVM), and (5) the dynamic equations of a
slider–crank mechanism for the simulations and comparisons.

2.1. Reynolds equation

The basis of the hydrodynamic lubrication theory is the Reynolds equation, which gives the pressure distribution in the
lubricating fluid. The classical form of the isothermal Reynolds equation for a dynamically loaded journal bearing can be
expressed as [2]:
   
∂ ∂p ∂ ∂p ∂h ∂h
h3 þ h3 ¼ 6μU þ 12μ ð1Þ
∂x ∂x ∂z ∂z ∂x ∂t

where x is the circumferential direction, z is the axial direction, μ is the fluid viscosity, h denotes the oil film thickness,
p is the pressure, and U represents the relative tangential velocity between journal and bearing surfaces. The two terms
on the right-hand side of Eq. (1) represent the two different effects of pressure generation on the lubricant film, the
wedge and the squeeze actions, respectively. The basic geometric parameters of a typical cylindrical journal bearing is
shown in Fig. 1.

Fig. 3. Sommerfeld and Gümbel boundary conditions (adapted from Machado et al. [2]).
G.B. Daniel et al. / Mechanism and Machine Theory 99 (2016) 19–36 23

The film thickness, h, according to the bearing geometry, can be express as

h ¼ C r þ e  cosðθÞ ¼ C r  ð1 þ ε  cosðθÞÞ ð2Þ

where Cr is the radial clearance, θ is the angular coordinate, e is the eccentricity, and ε = e/Cr is the eccentricity ratio.
Regarding the two terms on the right-hand side of Eq. (1), the first term represents the wedge effect (∂ h/∂ x), while the
second term represents the squeeze effect (∂ h/∂ t) of the oil film. The velocity component U can be obtained from the velocities
_ and Ω, as shown in Fig. 2, and the squeeze effect can be obtained deriving the film thickness (Eq. (2)) with respect
_ ϕ,
e,
to time [3]:

U ¼ C r  ε_  sinðθÞ − C r  ε  ϕ_  cosðθÞ þ Ω  R
ð3Þ
∂h=∂t ¼ C r  ε  ϕ_  sinðθÞ þ C r  ε_  cosðθÞ

where R represents the journal bearing radius and Ω is the relative angular velocity between the journal and the bearing. The
angle defined by the eccentricity vector is denoted by ϕ (Fig. 1). The dot on the top of any parameter in Eq. (3) denotes the
time derivative of such parameter.
Using the expressions in Eqs. (2) and (3), solving the corresponding derivatives and writing Eq. (1) in cylindrical coordinates
(x = Rθ), the Reynolds equation can be rewritten as

       
1 ∂ ∂p ∂ ∂p
2 ∂θ
h3
þ h3
¼ 6  μ  C r  2  ε_  cosðθÞ þ ε  2  ϕ_ − Ω  sinðθÞ ð4Þ
R ∂θ ∂z ∂z

It is important to point out that, for obtaining Eq. (4), the terms C 2r /R are neglected regarding the magnitude of the
others terms.
The exact solution of the Reynolds equation is difficult to obtain and, in general, requires a considerable numerical
effort [3]. However, it is possible to solve the equation analytically by setting to zero either the first or the second
term on the left-hand side of Eq. (4). These solutions correspond to those for infinitely short and infinitely long journal
bearings, respectively.

2.2. Infinitely short journal bearing

Dubois and Ocvirk [24] considered a short journal bearing if the pressure gradient around the bearings circumference is
considerably smaller than the pressure gradient along its length. This assumption is valid for small length-to-diameter (L/D)
ratios. Hence, the Reynolds equation for an infinitely short journal bearing can be written as

     
∂ ∂p
h3 ¼ 6  μ  C r  2  ε_  cosðθÞ þ ε  2  ϕ_ − Ω  sinðθÞ ð5Þ
∂z ∂z

From Eq. (2), it can be seen that h is not dependent of the z coordinate; therefore, Eq. (5) becomes
       
_ 6  μ  2  ε_  cosðθÞ þ ε  2  ϕ_ − Ω  sinðθÞ
∂2 p 6  μ  C r  2  ε_  cosðθÞ þ ε  2  ϕ − Ω  sinðθÞ
¼ ¼ ð6Þ
∂z2 h3 C 2r  ð1 þ ε  cosðθÞÞ 3

Fig. 4. Boundaries conditions of Reynolds equation for the proposed approach.


24 G.B. Daniel et al. / Mechanism and Machine Theory 99 (2016) 19–36

Eq. (6) can be easily integrate using the correct boundary conditions, in order to obtain the pressure distribution generated in
the oil film. Thus, when the relative pressure is set to zero at journal bearing edges, in the z direction, the fluid film pressure can
be expressed in the following form [1]:
    !
3  μ  2  ε_  cosðθÞ þ ε  2  ϕ_ − Ω  sinðθÞ 2 L
2
pðθ; zÞ ¼ z − ð7Þ
C 2r  ð1 þ ε  cosðθÞÞ 3 4

In most of the problems related to multibody systems, it is convenient to evaluate the force components of the resultant
pressure in directions tangent and radial to the centers line (radial and tangential directions in Fig. 1). These force components
can be obtained by integrating the pressure field either in the entire domain 2π or half-domain π, as it is illustrated in Fig. 3
(adapted from Machado et al. [2]).
These boundary conditions, associated with the pressure field, correspond to Sommerfeld's and Gümbel's boundary conditions. In the
second case, the pressure field is integrated only over the positive part by setting the pressure in the remaining portion equal to zero.
Thus, for the Sommerfeld's conditions the force components of the fluid film for infinitely short journal bearing are written as [3]:
 
π  μ  L  R ε_  1 þ 2  ε
3 2

Fr ¼   5 ð8Þ
C 2r 1 − ε2 2

Analysis
Conditions

(a)

(b) (c)

Fig. 5. Comparison of the pressure distribution obtained from numerical solution and proposed approach. (a) Pin's orbit with analysis conditions; (b) low
eccentricity; (c) high eccentricity.
G.B. Daniel et al. / Mechanism and Machine Theory 99 (2016) 19–36 25

 
_
π  μ  L  R ε  Ω − 2  ϕ
3
Ft ¼   3 ð9Þ
C 2r 2  1 − ε2 2

where Fr is the radial component of the force and Ft is the tangential component. This situation corresponds to the Frêne et al. [3]
full film short journal bearing solution. In a similar way, for the Gümbel's conditions, the force components of the fluid film for
infinitely short journal bearing are written as [3]:
2   3
−μ L R
3 π  ε_  1 þ 2  ε2  
Fr ¼  2  4 pffiffiffiffiffiffiffiffiffiffiffiffi _
þ2ε  Ω−2ϕ 5
2
ð10Þ
2  C 2r  1 − ε2 1−ε2

μ  L3  R  ε h π   pffiffiffiffiffiffiffiffiffiffiffiffiffiffii
Ft ¼    4  ε_ þ  Ω − 2  ϕ_  1 − ε 2
2 2
ð11Þ
2  Cr  1 − ε
2 2

Eqs. (10) and (11) represent the Frêne et al. [3] half film short journal bearing solution.
These force components for the tangent and radial directions to the centers line, projected in the X and Y directions, shown in
Fig. 1, are given by

F X ¼ F r  cosðϕÞ − F t  sinðϕÞ ð12Þ

F Y ¼ F r  sinðϕÞ þ F t  cosðϕÞ ð13Þ

2.3. Proposed approach to estimate the hydrodynamic forces for infinitely short bearings

As discussed in the first section of this paper, the main difficulty in obtaining the solutions of the hydrodynamic forces of
journal bearings lies not only in solving the differential equations but also in adequately defining the boundary conditions of
the Reynolds equation. The integration of the Reynolds equation only over the positive pressure regions, by assuming null the
pressure in the remaining portions, involves finding the zero points, i.e., the angle for which a positive pressure begins and the
angle for which the pressure is null. In the early 80s and 90s, some authors assumed these angles were generally equal to 0
and π for the case of a steady-state journal bearing, based on Gümbel model [3,25], although it is well known that these angles
can assume others values. However, for a dynamically loaded journal bearing, these angles are time dependent, and the evaluation
of the force components involves a lot of mathematical manipulation [2].
In this section, it is presented a method for the evaluation of the hydrodynamic force components obtained from the integra-
tion of the Reynolds equation for infinitely short bearings. Differently of the boundary conditions proposed by Sommerfeld and by
Gümbel [3], this work deals with an approach to determine the integration boundaries, i.e., the angular position in the bearing

Fig. 6. Scheme of the slider–crank mechanism. (A) Planar view. (B) Planar view with the forces over each component.
26 G.B. Daniel et al. / Mechanism and Machine Theory 99 (2016) 19–36

where the pressure is null. Therefore, the hydrodynamic forces are calculated taking into account only the positive pressure. As
previously mentioned, these angles are time dependent, and its evaluation is very important for a more accurate estimation of
the hydrodynamic force components.
The analysis starts from the expression of the pressure distribution generated in the oil film for the consideration of infinitely
short journal bearing, presented in Eq. (7) and reproduced below:

    !
3  μ  2  ε_  cosðθÞ þ ε  2  ϕ_ − Ω  sinðθÞ 2 L2
pðθ; zÞ ¼ z − ð7Þ
C 2r  ð1 þ ε  cosðθÞÞ 3 4

From Eq. (7), the angle for which the pressure is null, at the mid-length of the bearing (z = 0), can be obtained by imposing
that the numerator of this expression should be null. Then

   
2  ε_  cosðθÞ þ ε  2  ϕ_ − Ω  sinðθÞ ¼ 0 ð14Þ
2  ε_
tanðθÞ ¼   ð15Þ
ε  Ω − 2  ϕ_

Therefore, the angle for which a positive pressure starts and the angle for which the pressure is null (the boundaries
conditions of Reynolds equation) can be defined as follows:

0 1
−1 @ 2  ε_
θ1 ¼ tan  A
ε  Ω − 2  ϕ_ ð16Þ

θ2 ¼ θ1  π

The ± signals to θ2 (Eq. (16)) are related to the conditions shown in Fig. 4. In the first condition (+ signal), the positive
pressure is in the region delimited by θ1 b θ b θ2 (continuous line); in the second condition (− signal), the positive pressure is
in the region between θ b θ1 and θ N θ2(dotted line). Thereby, after the θ1 calculation, it is important to check the pressure signal
in order to properly evaluate θ2.
It is important to highlight the approach proposed here only shift the pressure distribution regarding the local reference (hmax)
at the bearing, being its angular extension of 180° of the positive pressure. Although the numerical solution of Reynolds equation
can provide an angular extension higher than 180°, this variation is not expressive, mainly in low L/D ratios. In this work, this
variation can be observed only for the ratio L/D = 1.0.
In order to clarify the effect of the angular extension, the Reynolds equation was solved with both methods (numerical and
proposed) under specific operational conditions, namely, rotational speed, eccentricity velocity (ε_ x ,ε_ y), and eccentricity (εx,εy), in-
dependently of the system dynamics. The results obtained in these simulations are presented in Fig. 5, considering the bearing
with ratio L/D = 1.0.
Fig. 5a shows the pin's orbit obtained from the proposed model. Considering this orbit as reference, two points
(eccentricity ratios of 0.23 and 0.67) are used as conditions for the evaluation of the pressure distributions. Fig. 5b and c
shows the pressure distributions obtained from the numerical solution and proposed approach for both cases of eccentricity.
As can be observed, the variation of angular extension is not significant even for the extreme situation (L/D = 1.0) considered in
this work.

Table 1
Geometrical parameters considered for the slider–crank mechanism.

Parameters Values

Crank length 0.0508 m


Connecting rod length 0.2032 m
Crank inertia 0.006 kg/m
Connecting rod inertia 0.010 kg/m
Crank mass 0.80 kg
Connecting rod mass 1.36 kg
Slider mass 0.91 kg
Position of the crank center of mass in the axis UM 0m
Position of the crank center of mass in the axis VM 0m
Position of the connecting rod center of mass in the axis UB 0.0508 m
Position of the connecting rod center of mass in the axis VB 0m
G.B. Daniel et al. / Mechanism and Machine Theory 99 (2016) 19–36 27

Table 2
Geometrical parameters considered for the journal bearing.

Parameters Values

Bearing's diameter 0.020 m


Radial clearance 20.10−6 m
Absolute viscosity 0.0117 Pa⋅s
Length–diameter ratios (L/D) 0.2; 0.3; 0.4; 0.5; 1.0

After defining the correct values for the integration limits, the expression for the pressure distribution, Eq. (7), can be integrated to
obtain the hydrodynamic forces components:
Z θ Z L=2
2
Fr ¼ p  cosðθÞ  dz  Rdθ ð17Þ
θ1 −L=2

Z θ2
Z L=2
Ft ¼ p  sinðθÞ  dz  Rdθ ð18Þ
θ1 −L=2

Using Eq. (7), Eqs. (17) and (18) can be developed as follows:
" Z Z #
μ RL
3 θ2 2
cos ðθÞ   θ2 sinðθÞ  cosðθÞ
Fr ¼ − 2  ε_  dθ þ ε  2  ϕ_ − Ω  dθ ð19Þ
2 θ1 ð1 þ ε  cosðθÞÞ3
θ1 ð1 þ ε  cosðθÞÞ3

" Z Z #
μ RL
3 θ2 sinðθÞ  cosðθÞ   θ2 2
sin ðθÞ
Ft ¼ − 2  ε_  dθ þ ε  2  ϕ_ − Ω  dθ ð20Þ
2 θ1 ð1 þ ε  cosðθÞÞ3
θ1 ð1 þ ε  cosðθÞÞ3

The resulting integrals in θ are solved using the bound values θ1 and θ2 evaluate by Eq. (16).
In this work, the integrals (reproduced in Eq. (21)) are solved numerically using the limits given by Eq. (16), in order to obtain
the correct values for the hydrodynamic forces components for the case of infinitely short bearings.
Regarding the numerical integration, it is used as a class of quadrature rules that can be constructed by interpolating functions
that are easy to integrate. Typically, these interpolating functions are polynomials. As the intention of this paper is not to evaluate
the capabilities of the methods, it is used as the simplest method of this type, which consists in letting the interpolating function
to be a constant function, a polynomial of zero degree. This is called the rectangle rule:
Z θ2 cos2 ðθÞ

θ1 ð1 þ ε  cosðθÞÞ3
Z θ2 sinðθÞ  cosðθÞ
dθ ð21Þ
θ1 ð1 þ ε  cosðθÞÞ3
Z θ2 sin2 ðθÞ

θ1 ð1 þ ε  cosðθÞÞ3

(a) (b)
Fig. 7. Comparison using different solutions for ratio of 0.2. (a) Oil film thickness. (b) Pin's orbit.
28 G.B. Daniel et al. / Mechanism and Machine Theory 99 (2016) 19–36

Fig. 8. Comparison using different solutions for ratio of 0.2. (a) Force in X direction. (b) Force in Y direction.

The hydrodynamic force components are hence projected in the x and y directions using Eqs. (12) and (13).
It is well known that the boundary conditions applied to the Reynolds equation solution, and consequently, the cavita-
tion model, depends on the oil supply to the bearing. Hence, the point here is that the oil supply must, in this case, be
necessary and sufficient to guarantee the full immersion inside the bearing (flooded bearing), for all models discussed in
this work.

2.4. Numerical solution of the Reynolds equation

The numerical solution of the Reynolds equation is obtained from the FVM, where the pressure in a specific volume (PP) is
obtained using the adjacent volumes (E: eastern; W: western; N: northern; S: southern) and the source term (B). Thus, the
pressure distribution can be calculated as shown in Eq. (22):

CP PP ¼ CE PE þ C W PW þ CN PN þ CS PS þ B ð22Þ

where the coefficients in Eq.(22) are determined as


! ! ! !
Δz h3e Δz h3w Δx h3n Δx h3s
CE ¼ ; CW ¼ ; CN ¼ ; CS ¼ ;
Δx μ Δx μ Δz μ Δz μ ð23Þ
∂h
CP ¼ CE þ CW þ CN þ CS ; B ¼ − 6U ½he − hw  − 12 ΔxΔz
∂t

The hydrodynamic forces can be obtained by the numerical integration of the pressure distribution. Eq. (24) presents
the hydrodynamic forces components in the radial and tangential directions, where Nx and NZ are respectively the total
number of points in the mesh in each direction. Finally, these forces are projected in the x and y directions also using
Eqs. (12) and (13):

8 9
>
> X
Nx X
NZ >
>
>
> P P sinðθÞ ΔxΔz >
>
>
< >
=
Fr i¼1 j¼1
¼ ð24Þ
Ft >
> X
Nx X
NZ >
>
>
> − P P cosðθÞ ΔxΔz >
>
>
: >
;
i¼1 j¼1

Table 3
Global error for ratio of 0.2.

Lubrication models Global error (%)

εx εy Fx Fy

Gümbel 71.4 102.8 91.7 98.1


Sommerfeld 96.1 95.3 32.9 39.2
Proposed 0.6 14.4 14.9 15.3
G.B. Daniel et al. / Mechanism and Machine Theory 99 (2016) 19–36 29

(a) (b)
Fig. 9. Comparison using different solutions for ratio of 0.3. (a) Oil film thickness. (b) Pin's orbit.

2.5. Dynamic equations of the mechanism

The mechanical system considered in the simulations and, consequently, the comparisons shown in this paper concern to the slider–
crank mechanism used by Daniel and Cavalca [15]. The classical slider–crank mechanism has one degree of freedom and a planar motion.
The consideration of a clearance in the piston-pin joint (journal bearing) does not change the planar motion characteristic; however, it
adds two degrees of freedom to the system. The chosen independent coordinates are the crank angular displacement (q), the connecting
rod angular displacement (A), and the slider displacement (XPT). Fig. 6 shows a scheme of this mechanism.
Defining IMO and IB as the crank and connecting rod rotational inertias and applying Lagrange's formulation, the equation of
motion for the complete mechanism is given by Eq. (25) [15]:
" 2
#
MB R þ IM0 − MB R ðU PB cos
 ðA þ qÞ þV PB sinðA þ qÞÞ €
q
sym
2 2
MB U PB þ V PB þ IB € þ
A
A_
2
þMB R ðU PB sinðA þ qÞ − V PB cosðA þ qÞÞ
_q2
F yp R cosðqÞ − F xp R sinðqÞ − τd  
MM ð − U PM sinðqÞ − V PM cosðqÞ − MB R sinðqÞÞ € ¼ − F þ gM þ F
þg ¼ MPT X
MB ð−U PB sinðAÞ þ V PB cosðAÞÞ − F yp L cosðAÞ − F xp L sinðAÞ PT ext PT xp

ð25Þ

(a) (b)

Fig. 10. Comparison using different solutions for ratio of 0.3. (a) Force in X direction. (b) Force in Y direction.
30 G.B. Daniel et al. / Mechanism and Machine Theory 99 (2016) 19–36

(a) (b)
Fig. 11. Comparison using different solutions for ratio of 0.4. (a) Oil film thickness. (b) Pin's orbit.

where g is the gravitational acceleration; MM is the crank mass; MB is the connecting rod mass; MPT is the slider mass; UPM and
VPM are the coordinates of the crank center of mass in the local reference system (UM, VM); UPB and VPB are the coordinates of the
connecting rod center of mass in the local reference system (UB, VB); XPM and YPM are the displacements of the crank's center of
mass; XPB and YPB are the displacements of the connecting rod's center of mass; and XPT is the displacement of the piston, in the
inertial reference system (X, Y).

3. Results and discussion

In this work, the computational simulations aim to compare different analytical solutions with the numerical solution of the
Reynolds equation, allowing to evaluate the representativeness of the analytical models (closed solutions) for short bearings. The
mechanical system considered in the simulations is the slider–crank mechanism used by Daniel and Cavalca [15] and described in
Section 2.5. The geometric parameters of the mechanism and of the bearing are presented in Tables 1 and 2, respectively.
The computational simulations consider the slider initially located at the top dead center and the crank's initial angular
velocity of 250 rad/s, as initial conditions. All results plots are taken after 4 laps of the crank in order to guarantee the steady
state of the system response. The equations of motion are time-integrated, using a predictor–corrector integrator proposed by
Shampine and Gordon [26], largely applied in dynamic simulations of nonlinear systems. This integrator works with a variable
time-step and with an error control.

Table 4
Global error for ratio of 0.3.

Lubrication models Global error (%)

εx εy Fx Fy

Gümbel 71.4 149.4 26.8 34.3


Sommerfeld 98.3 101.8 15.3 17.4
Proposed 1.7 11.2 8.8 9.1

Table 5
Global error for ratio of 0.4.

Lubrication models Global error (%)

εx εy Fx Fy

Gümbel 72.0 188.0 12.6 16.9


Sommerfeld 99.5 110.1 8.8 9.5
Proposed 3.2 10.5 6.2 5.9
G.B. Daniel et al. / Mechanism and Machine Theory 99 (2016) 19–36 31

Table 6
Global error for ratio of 0.5.

Lubrication models Global error (%)

εx εy Fx Fy

Gümbel 71.8 215.1 7.9 10.4


Sommerfeld 99.0 135.0 6.2 6.9
Proposed 5.0 12.0 4.5 4.0

The numerical solution, developed from FVM (Section 2.4), used the computational mesh with 900 volumes (30 volumes per
direction x and z). It was verified that this amount of volumes was able to satisfactory converge the pressure distribution for all
cases with different L/D ratios. Regarding the proposed solution, the numerical integration (Eq. (21)) is calculated from the
rectangle rule, using 100 points, sufficient amount for a satisfactory approximation of the integrals.
As verified in Table 2, the comparison of the analytical and numerical solutions is carried on for different length–diameter
ratios of the bearing, in order to verify the L/D range with good agreement to the solution for finite bearing (numerical solution).

(a) (b)
Fig. 12. Comparison using different solutions for ratio of 0.4. (a) Force in X direction. (b) Force in Y direction.

(a) (b)
Fig. 13. Comparison using different solutions for ratio of 0.5. (a) Oil film thickness. (b) Pin's orbit.
32 G.B. Daniel et al. / Mechanism and Machine Theory 99 (2016) 19–36

(a) (b)
Fig. 14. Comparison using different solutions for ratio of 0.5. (a) Force in X direction. (b) Force in Y direction.

The comparison of the numerical solution and the others models is carried on evaluating a global error as shown in
Eq. (26):

NP qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
X NP qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
X
ð F models − F numerical Þ2 ðDmodels −Dnumerical Þ2
i¼1 i¼1
NP NP
Δforce ¼ N P qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Δdisplacements ¼ NP qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ð26Þ
X X
ð F numerical Þ2 ðDnumerical Þ2
i¼1 i¼1
NP NP

where Δforce and Δdisplacements are the global errors of the hydrodynamic forces and pin's displacements, respectively; Fnumerical
and Fmodels are the hydrodynamic forces obtained from the numerical solution and the others models, respectively; Dnumerical
and Dmodels are the pin's displacements obtained from the numerical solution and the others models, respectively; and NP is
the number of points in time domain.
Initially, the results obtained for the length–diameter ratio of 0.2 are presented. Fig. 7 shows the oil film thickness and the
pin's orbit inside the bearing, comparing the analytical closed solutions (Gümbel and Sommerfeld boundary conditions), the

(a) (b)
Fig. 15. Comparison using different solutions for ratio of 1.0. (a) Oil film thickness. (b) Pin's orbit.
G.B. Daniel et al. / Mechanism and Machine Theory 99 (2016) 19–36 33

(a) (b)
Fig. 16. Comparison using different solutions for ratio of 1.0. (a) Force in X direction. (b) Force in Y direction.

proposed, and the numerical solutions. As can be verified in Fig. 7, the load capacity of the bearing is low due to the length–diameter
ratio, and consequently, the pin's trajectory reaches values close to the internal surface of the bearing. Thus, the minimum oil film
thickness is very small, but there is no contact between pin and bearing. However, under these extremes eccentricities, it is
important to point out that the lubrication regime is probably an elastohydrodynamic lubrication, as analyzed in others work
[4,8,19,27]. The proposed approach does not take into account the mechanical deformation effects since the main focus here is the
evaluation of the proposed solution for hydrodynamic regime regarding other similar approaches (classical solutions).
Moreover, some important characteristics can be verified comparing the results obtained for the different solutions. The
analytical closed solutions provide results significantly different of the numerical solution. On the other hand, the results obtained
from the approach proposed here and the numerical solutions are in good agreement. This behavior is a consequence of the
adjustment of the boundary conditions for the integration of the pressure distribution in the bearing, differently from the half
film (Gümbel solution) and full film (Sommerfeld solution).
The hydrodynamic forces obtained for different solutions are shown in Fig. 8. The tendencies of the hydrodynamic forces are
similar for all solutions in both directions X and Y. However, the hydrodynamic forces obtained from the Gümbel solution presents
high numerical oscillations when the pin and the bearing are critically close. This behavior can be in consequence of the model
sensitivity (Gümbel solution) because at the higher eccentricity levels the system of equations becomes stiff. In fact, the other
solutions also present numerical oscillations when the pin and the bearing are critically close, but in significantly lower levels.
As can be seen in the details of Fig. 8, the results obtained from the proposed approach and the numerical solution are in good

(a) (b)
Fig. 17. Comparison using proposed solution and numerical solution for infinitely short bearing considering the ratio of 0.2. (a) Oil film thickness. (b) Pin's orbit.
34 G.B. Daniel et al. / Mechanism and Machine Theory 99 (2016) 19–36

(a) (b)
Fig. 18. Comparison using proposed solution and numerical solution for infinitely short bearing considering the ratio of 0.2. (a) Force in X direction. (b) Force
in Y direction.

agreement. Moreover, Table 3 shows that the global errors for the proposed approach are much smaller than those obtained with
the classical solutions, for both displacements and forces.
As previously mentioned, the same simulations are performed for the length–diameter ratios of 0.3, 0.4, and 0.5. Figs. 9, 11,
and 13 show the oil film thickness and the pin's orbit inside the bearing obtained for all the mentioned cases. The results in
these simulations are similar to the previous one, i.e., the analytical closed solutions provide results significantly different from
the numerical solution while the proposed approach is always in good agreement regarding the numerical solution. However,
the minimum oil film thickness is higher as the L/D ratio increases since the load capacity of the bearing is also higher. For
this reason, numerical oscillations usually present in the hydrodynamic forces are lower, mainly in the results obtained by Gümbel
solution, as can be verified in Figs. 10, 12 and 14. The hydrodynamic forces obtained for all cases and solutions have the same
tendencies in both directions X and Y. In fact, also in these cases a good agreement can be observed between the proposed
approach and the numerical solutions. Moreover, due to the alternate motion of the slider–crank mechanism, for all cases, the
hydrodynamic force in X direction is greater than in Y direction.
According to Figs. 7, 9, 11, and 13, the differences between the results obtained from the proposed approach and the numerical
solution tend to increase as the length–diameter ratio increases. This tendency can also be observed by analyzing the global errors
presented in Tables 4, 5, and 6.
In fact, this behavior is already expected since the numerical solution is obtained from the complete Reynolds equation while
the proposed approach is developed for infinitely short bearing. However, the differences are still very low, which indicate that
the proposed approach is promising to efficiently represent the hydrodynamic conditions in bearings with length–diameter
ratio up to 0.5.
In order to verify the differences between the solutions for finite and infinitely short bearings, the results for the computational
simulations considering the length–diameter ratio of 1.0 are shown in Figs. 15 and 16.
Fig. 15 shows the oil film thickness and the pin's orbit inside the bearing obtained from the different solutions. In this case,
oil film thicknesses present different values as for the analytical closed solutions as for the proposed approach, when the pin is
closer to the bearing wall. However, the pin's trajectories, for the proposed approach and the numerical solution, present a similar
tendency. Special attention must be taken under this condition since the proposed solution cannot efficiently represent the
lubrication condition in the bearing due to the high L/D ratio (not attending the infinitely short bearing assumption).
Regarding the hydrodynamic forces, no expressive difference can be verified in the results obtained for the length–diameter
ratio of 1.0, as shown in Fig. 16. In this case, the hydrodynamic forces present a similar tendency, and the numerical oscillations
are significantly reduced since the pin is not critically close of the bearing (small eccentricity/high minimum oil film thickness).

Table 7
Global error for ratio of 1.0.

Lubrication models Global error (%)

εx εy Fx Fy

Gümbel 67.6 233.7 3.1 4.5


Sommerfeld 78.2 652.0 2.3 7.4
Proposed 18.1 23.6 1.4 1.6
G.B. Daniel et al. / Mechanism and Machine Theory 99 (2016) 19–36 35

Table 8
Global error for ratio of 0.2 considering infinitely short bearing.

Lubrication models Global error (%)

εx εy Fx Fy

Proposed 0.1 5.23 12.9 12.4

As observed in Tables 3-8, the global error of the hydrodynamic forces decreases due to the lower numerical oscillations as
the L/D ratio increases. However, when looking at the pin trajectories (orbits), it is possible to observe that the proposed approach
is in better agreement with the numerical solution for lower L/D ratios. The displacements are less sensitive to the numerical
oscillations in the hydrodynamic forces than velocity and acceleration [27]. This is an important issue that defines how much
the lubrication proposed approach is close to the actual lubrication condition.
According to Figs. 15 and 16, the system dynamic behavior under different lubrication models leads to different dynamic
conditions for the pin, namely, eccentricity and eccentricity velocity, in order to generate similar hydrodynamic forces and,
consequently, to satisfy the equation of motion of the mechanical system. Therefore, the tendency of the hydrodynamic forces
is similar for different L/D ratios.
The last two figures, Figs. 17 and 18, present the comparison between the proposed solution and the numerical solution of the
Reynolds equation for the consideration of infinitely short bearing, i.e., the numerical solution of Eq. (5). The results are for the
length–diameter ratio of 0.2, the case for shortest bearing. Fig. 17 shows the oil film thickness and the pin's orbit inside the
bearing, and Fig .18 shows the hydrodynamic forces. The aim here is to show that the two solutions (numerical and proposed)
are very closed. The global errors obtained in this case occur due to the application of different solution methods for Eq. (5);
therefore, these errors are low as shown in Table 8.
According to Fig. 17, the oil film thickness and the pin's orbit are overlapped, indicating that both solutions are quite
coincident. Regarding the hydrodynamic forces, the results are also overlapped but not completely since oscillations are present
in the response when the oil film thickness is low, as can be seen in the detail of Fig. 18.
Finally, a comparison of the computational costs to the numerical simulations is accomplished. Table 9 shows that the
solutions proposed by Sommerfeld and by Gümbel are faster than the proposed approach. In fact, this behavior is quite expected
since these solutions are analytical closed solutions while the proposed approach involves a numerical integration (Eq. 21).
Despite the analytical closed solutions can be fast, the respective responses are not in a good correlation with the numerical
solution, compromising their applications in some critical cases. For this reason, the proposed method emerges as very promising
since it can be significantly faster than the numerical solution (in some cases more than 850 times faster) and provides a good
agreement of the results.
All simulations were performed in the same personal computer (desktop with Core i7-3770 K Processor 3.4 GHz 8 M Cache
and Memory/RAM 16GB DDD3). Therefore, it is possible to observe that the time spent on the computer simulation using the
proposed approach is quite acceptable, making this approach feasible for many applications.

4. Conclusion

The paper presents a discussion about the integration bounds of Reynolds equation for infinitely cylindrical short journal
bearings, i.e., the initial angular position in the bearing where the pressure is positive. The proposed approach evaluates the
pressure distribution under dynamic conditions, namely, taking into account the pin's trajectory in time domain, which is not
commonly addressed in the available literature in the field. Actually, the angular boundaries that limit the positive pressure
field from the null pressure are time dependent and have influence the nonlinear hydrodynamic forces. This fact directly affects
the orbital position of the shaft inside the bearing, the oil film thickness, and successively, the dynamics of the system. This
phenomenon occurs due to the squeeze film effect.
The comparison of the proposed model with the complete numerical solution of Reynolds equation is promising and in good
agreement, as observed in the pin's orbits and in the global error analysis. This similarity is not always true for the classical
models. Moreover, the contribution of the proposed approach to the computational costs is significant, being its absolute compu-
tational time spent in the simulations considerably lower than for numerical solution of the same case. Hence, the proposed

Table 9
Time [s] spent on computer simulations.

Length–diameter ratio (L/D)


Solution
0.2 0.3 0.4 0.5 1.0

Gümbel's boundary condition 77.5 61.8 58.7 59.0 58.0


Sommerfeld's boundary condition 63.0 60.0 58.0 60.2 58.0
Proposed model 81.9 76.2 65.0 63.1 59.7
Numerical solution 53906 26040 16365 10316 2971
36 G.B. Daniel et al. / Mechanism and Machine Theory 99 (2016) 19–36

approach is promising for applications in short bearings, giving good quality results in feasible short time, highly desirable in the
design process of the complete mechanical system.
Finally, the novelty of the approach proposed in this paper is to introduce the possibility to shift the pressure distribution
regarding the reference location (hmax) at the bearing. Consequently, this approach can be inserted into a closed model for the
adjustment of boundary conditions for the use in simulations of mechanisms and mechanical systems in time domain.

Acknowledgements

The authors thank FAPESP, CAPES, and CNPq for the financial support of this research.

References

[1] O. Pinkus, S.A. Sternlicht, Theory of Hydrodynamic Lubrication, McGraw-Hill, New York, 1961.
[2] M. Machado, J. Costa, E. Seabra, P. Flores, The effect of the lubricated revolute joint parameters and hydrodynamic force models on the dynamic response of
planar multibody systems, Nonlinear Dyn. 69 (2012) 635–654.
[3] J. Frêne, D. Nicolas, B. Degneurce, D. Berthe, M. Godet, Hydrodynamic lubrication, Bearings and Thrust Bearings, Elsevier, Amsterdam, 1997.
[4] P. Flores, J. Ambrósio, J.P. Claro, Dynamic analysis for planar multibody mechanical systems with lubricated joints, Multibody Sys. Dyn. 12 (2004) 47–74.
[5] P. Flores, H.M. Lankarani, J. Ambrósio, J.C.P. Claro, Modelling lubricated revolute joints in multibody mechanical systems, Proc. Inst. Mech. Eng. K J. Multi-body
Dyn. 218 (4) (2004) 183–190.
[6] B.J. Alshaer, H. Nagarajan, H.K. Beheshti, H.M. Lankarani, S. Shivaswamy, Dynamics of a multibody mechanical system with lubricated long journal bearings,
J. Mech. Des. 127 (2005) 493–498.
[7] P. Flores, J. Ambrósio, J.C.P. Claro, H.M. Lankarani, C. Koshy, A study on dynamics of mechanical systems including joints with clearance and lubrication, Mech.
Mach. Theory 41 (2006) 247–261.
[8] P. Flores, J. Ambrósio, J.C.P. Claro, H.M. Lankarani, C.S. Koshy, Lubricated revolute joints in rigid multibody systems, Nonlinear Dyn. 56 (2009) 277–295.
[9] Q. Tian, Y. Zhang, L. Chen, J. Yang, Simulation of planar flexible multibody systems with clearance and lubricated revolute joints, Nonlinear Dyn. 60
(2010) 489–511.
[10] P. Ravn, S. Shivaswamy, B.J. Alshaer, H.M. Lankarani, Joint clearances with lubricated long bearings in multibody mechanical systems, J. Mech. Des. 122
(2000) 484–488.
[11] J.F. Booker, Dynamically loaded journal bearings: mobility method of solution, Trans. ASME J. Basic Eng. 4 (1965) 537–546.
[12] J.F. Booker, Dynamically loaded journal bearings: numerical application of mobility method, Trans. ASME J. Lubr. Technol. 1 (1971) 168–176.
[13] P. Flores, J.C.P. Claro, J. Ambrósio, Journal bearings subjected to dynamic loads: the analytical mobility method, Mec. Exp. Assoc. Port. Anál. Exp. Tensões 13
(2006) 115–127.
[14] J.-H. Choi, S.J. Lee, D.-H. Choi, Tolerance optimization for mechanisms with lubricated joints, Multibody Sys. Dyn. 2 (2) (1998) 145–168.
[15] G.B. Daniel, K.L. Cavalca, Analysis of the dynamics of a slider–crank mechanism with hydrodynamic lubrication in the connecting rod-slider joint clearance, Mech.
Mach. Theory 46 (2011) 1434–1452.
[16] A.C. Bannwart, K.L. Cavalca, G.B. Daniel, Hydrodynamic bearings modeling with alternate motion, Mech. Res. Commun. 37 (2010) 590–597.
[17] D. Wang, T.G. Keith, Q. Yang, K. Vaidyanathan, Lubrication analysis of a connecting-rod bearing in a high-speed engine. Part I: rod and bearing deformation, Tribol.
Trans. 47 (2004) 280–289.
[18] D. Wang, T.G. Keith, Q. Yang, K. Vaidyanathan, Lubrication analysis of a connecting-rod bearing in a high-speed engine. Part II: lubrication performance evaluation
for non-circular bearings, Tribol. Trans. 47 (2004) 290–298.
[19] Q. Tian, Y. Sun, C. Liu, H. Hu, P. Flores, ElastoHydroDynamic lubricated cylindrical joints for rigid-flexible multibody dynamics, Comput. Struct. 114–115
(2013) 106–120.
[20] Q. Tian, Q. Xiao, Y. Sun, H. Hu, H. Liu, P. Flores, Coupling dynamics of a geared multibody system supported by ElastoHydroDynamic lubricated cylindrical joints,
Multibody Sys. Dyn. 33 (2015) 259–284.
[21] A.L. Schwab, J.P. Meijaard, P. Meijers, A comparison of revolute joint clearance model in the dynamic analysis of rigid and elastic mechanical systems, Mech. Mach.
Theory 37 (9) (2002) 895–913.
[22] S. Erkaya, S. Doğan, S. Ulus, Effects of joint clearance on the dynamics of a partly compliant mechanism: numerical and experimental studies, Mech. Mach. Theory
88 (2015) 125–140.
[23] P. Flores, H.M. Lankarani, Spatial rigid-multi-body systems with lubricated spherical clearance joints: modeling and simulation, Nonlinear Dyn. 60 (1–2)
(2010) 99–114.
[24] G.B. Dubois, F.W. Ocvirk, Analytical derivation and experimental evaluation of short bearing approximation for full journal bearings, NACA Rep. 1157 (1953).
[25] G. Capone, Orbital motions of rigid symmetric rotor supported on journal bearings, Mec. Ital. 199 (1986) 37–46 (n.).
[26] L.F. Shampine, M.K. Gordon, Computer Solution of Ordinary Differential Equations: The Initial Value Problem, W.H.Freeman & Co Ltd, 1975.
[27] V.L. Reis, G.B. Daniel, K.L. Cavalca, Dynamic analysis of a lubricated planar slider–crank mechanism considering friction and Hertz contact effects, Mech. Mach.
Theory 74 (2014) 257–273.

You might also like