Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Numerical Study of the Lubricant

Viscosity Grade Influence on Thrust


Bearing Operation

Nikolay Sokolov, Mullagali Khadiev, Pavel Fedotov , and Eugeny Fedotov

Abstract The article presents studies of the influence of the viscosity grade of the
supplied oil ISO VG32 and VG46 in a wide range of rotor speeds and operating clear-
ances on the local and integral characteristics of a thrust plain bearing with fixed pads
of the compressor. The studies were carried out using the Sm2Px3Txτ calculation
program based on the results of numerical experiments of the bearing. The program is
built by numerical implementation of a non-stationary periodic thermoelastohydro-
dynamic (PTEHD) mathematical model of the thrust bearing operation. The research
results indicate a significant influence of the oil viscosity grade on the main charac-
teristics and temperature conditions of the thrust bearing. When changing from ISO
VG46 to the lighter oil VG32, there is a noticeable reduction in bearing pad tempera-
tures and power loss. However, the level of this change is determined by the specified
operating clearance between the rotating collar and the bearing pads. The influence
of oil viscosity grade and the profile of the active surface on the temperature regime
of the pad is analyzed. The value and location of the maximum temperature of the
thrust bearing pad is determined, as well as the possibility of applying the standard
point 75/75 from API-670 in practice.

Keywords Thrust bearing · Mathematical model · Lubricating and boundary


films · Fixed pad · Numerical method · Boundary value problem · Viscosity
grade · Circumferential velocity · Clearance height

N. Sokolov (B) · M. Khadiev


Kazan National Research Technological University, 10 Popov, Kazan 420029, Russia
e-mail: sokol-88@list.ru
P. Fedotov
Kazan Federal University, 35 Kremlevskaya, Kazan 420008, Russia
E. Fedotov
AST Volga Region LLC, 50 Petersburgskaya, Kazan 420107, Russia

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 205
A. G. Kravets et al. (eds.), Cyber-Physical Systems Engineering and Control,
Studies in Systems, Decision and Control 477,
https://doi.org/10.1007/978-3-031-33159-6_16
206 N. Sokolov et al.

1 Introduction

The hydrodynamic fixed-geometry thrust bearing of a centrifugal and screw


compressor is a critical element of the casing design, the quality of which deter-
mines the reliability of the compressor as a whole. The temperature regime of the
bearing is largely determined by the operating parameters (external factors affecting
it), which include the rotor speed, temperature and viscosity of the supplied lubri-
cant, lubricant consumption, axial load from the pressure drop of the compressible
gas, as well as the nature of the change in this load over time. During the operation
of a thrust bearing, diagnostic signals are needed to monitor the state of its opera-
tion. The generally accepted criteria for performance under high heat conditions, and
therefore the limiting parameters for plain bearings, are the minimum thickness and
maximum temperature of the lubricating film. The main diagnostic signal, despite
the large temporal inertia, remains the temperature of the active surface of the pad
[1]. For example, a chromel–alumel or chromel-kopel thermocouple in the solid of
the pad near the babbitt layer of the bearing. In some cases, due to the difficulty of
removing the airbag signal cable from the housing and due to frequent breaks during
assembly and disassembly, the temperature of the lubricant at the drain from the
bearing housing is used, despite incomplete information [2].
The limiting temperatures of the lubricating film of a plain bearing at increased
circumferential velocities and loads can have a significant effect on the mechanical
properties of the anti-friction coating of the pads. In the case of using a babbit coating,
its fatigue strength is significantly reduced, which leads over time, under the influence
of a periodic external load, to the appearance of cracks on the active surface of the
pad and further chipping of the babbit [3]. Bearings with babbitt coating B-83 or
Tegostar 738 (contains 81.3% tin Sn), which are most often used in compressors,
are recommended for use at lubrication temperatures not exceeding 115 °C [3]. With
an increase in the tin content, the allowable operating temperature of the material
decreases. It should also be taken into account that increased temperature operating
conditions of bearings lead to aging and the appearance of coke-like products in the
oil over time [4]. In accordance with this, according to [5], the limiting temperature
for mineral oils most used in compressor engineering is a temperature of 120–140 °C.
The limiting temperature for B-83 babbitt is 110–115 °C [5]. Thus, for the journal
and thrust bearings of the compressor, the allowable operating temperature can be
considered to be [tmax ] = 110 °C [3].
Lubrication is supplied to the thrust bearing to ensure the formation of a hydro-
dynamic film and the removal of heat generated during operation due to the viscous
shear of the lubricant layers. At high shear rates, this heat dissipation (or power loss)
can be significant in terms of overall compressor power input and energy efficiency.
In connection with the study of the general temperature regime of operation of the
thrust bearing, of interest are works that consider the change in the temperature
field of the bearing when the operating parameters change. Such works include, for
example, experimental studies [6–8], which studied the effect of the viscosity grade
of the supplied oil on the temperature distribution of tilting pads and the rotating
Numerical Study of the Lubricant Viscosity Grade Influence on Thrust … 207

element (collar or trunnion of the rotor), on the change in oil inlet and outlet temper-
atures, flow rates and eccentricity depending on the applied load and circumferential
velocity. The authors themselves confirm that the obtained isotherms of the active
surface of the pads in a wide range of speeds and loads were obtained by extrapolation
from the readings of thermocouples and, therefore, have inaccuracies.
To avoid such inaccuracies when analyzing the operation of the hydrodynamic
fixed-geometry thrust bearing the Sm2Px3Txτ program [9] allows. The program is
based on the numerical implementation of a non-stationary periodic thermoelasto-
hydrodynamic (PTEHD) model of operation of a thrust bearing. This formulation
allows us to consider jointly occurring hydrodynamic and thermal processes in the
lubricant film and solid elements of the bearing. The advantage of the PTEHD math-
ematical model is the description of a more complete thermal picture of the bearing
and, consequently, obtaining more reliable output data [10]: local characteristics
(maximum temperature and minimum thickness of the bearing lubricating film) and
integral characteristics (load-bearing capacity, frictional power losses, lubrication
costs through the inlet and outlet sections of the lubricating film, heat flows through
the sections of structural elements and the lubricant film, etc.), as well as the distri-
bution of pressure and temperatures in the lubricating and boundary films and solids,
including depending on time.
A feature of the PTEHD theory when setting periodic thermal boundary conditions
is the direct calculation of the lubricant temperature distribution in the inlet section
of the lubricating film, taking into account the mixing of two lubricant flows: a
heated flow from the previous pad, which moves under the action of a rotating collar
(sticking condition), and a cold lubrication from the inter-pad groove (IPG). Such a
setting also makes it possible to calculate the temperatures of the pad and the rotating
collar, taking into account heat exchange with the external environment. The latter
is especially important, since the collar acts as a heat accumulator, participating in
the redistribution of heat. When passing through the areas of the lubricating and
boundary films in the process of gradual heating of the lubricant in the direction of
rotation, the heat flow changes its direction from the collar to the lubricating film
and back [11]. The temperature of the rotating collar in this case assumes a certain
averaged constant value.
This article presents a description of some differential equations of the PTEHD
mathematical model, the basis of numerical implementation, and describes the results
of numerical experiments of the hydrodynamic fixed-geometry thrust bearing of a
centrifugal or screw compressor: the effect of oil supply with different viscosity
grade (ISO VG32 and VG46) on the general temperature state and bearing power
loss at different circumferential velocities and height of the working clearance.
208 N. Sokolov et al.

2 Description of the Mathematical Model

The developed periodic thermoelastohydrodynamic (PTEHD) mathematical model


is based on the fundamental laws of conservation of mass, momentum and energy,
which ensures the adequacy of the formulation and accuracy of calculations. Two
profiles of the active surface of pads are considered as the most used in centrifugal
and screw compressors: parallel taper land (Fig. 1a) and taper land (Fig. 1b) with a flat
land. The lubricating area of L2 the bearing is divided into a converging bearing lubri-
cating film of the pad and an expanding boundary film of the IPG. The boundary film
between the pads is determined by a conditional boundary, which allows geometri-
cally smoothly changing its thickness. In order to get acquainted with the description
of the mathematical model in the PTEHD formulation in relation to the operation of
a thrust bearing with fixed pads and some physics of the hydrodynamic process, one
can refer to the articles [12, 13].
The PTEHD mathematical model is presented in dimensional and dimension-
less forms through relative (i.e., dimensionless) quantities related to characteristic

a) b)

c)

Fig. 1 Calculation scheme of the thrust bearing: a pad profile with a parallel taper land; b pad
profile with a taper land; c section along A–A along the middle radius: 1—bevel part, 2—flat part
of the pad; 3—rotating collar; 4—lubricating film; 5—boundary film
Numerical Study of the Lubricant Viscosity Grade Influence on Thrust … 209

dimensions (the «-» sign above the quantity). The dimensionless view is necessary
to solve the equations by numerical methods and to simplify the parametric analysis
of the thrust bearing. The origin of the coordinates of the lubricating and boundary
films is located on the surface of the collar. The main governing equations with the
corresponding boundary conditions are:
1 the generalized Reynolds equation describing the two-dimensional distribution of
pressure in the lubricating film of the L 1 (−1 ≤ r ≤ 1, 0 ≤ ϕ ≤ θ p , 0 ≤ y ≤ 1)
area. The equation is derived with a minimum of restrictive assumptions, in which
the density and viscosity of the lubricant are functions of all three coordinates
(the form of the equation is close to the derivation by the author Dowson D. for
the Cartesian coordinate system [14]). In dimensionless non-stationary form, the
equation takes the following form
[ ] [ 3
]
∂ 3 ∂ p̄ ∂ h̄ ∂ p̄
−λ 2
(σ r̄ + 1)h̄ f̄ 0 − f̄
∂ r̄ ∂ r̄ ∂ ϕ̄ (σ r̄ + 1) 0 ∂ ϕ̄
( 3 ) (1)
( )
∂ h̄ f̄ 1 t ∂ h̄ f̄ 2
= −Reψσ λ 2
+ ω̄(σ r̄ + 1) + Sh(σ r̄ + 1) Ā,
∂ r̄ ∂ ϕ̄

( )
where r , ϕ, y are dimensionless coordinates;
( ) p = ph 2
20 / μ0 ω∗ R 2
av θ is the local
∫1
dimensionless pressure; A = ∂τ∂ h ρd y − ρ y=1 ∂h ∂τ
is a non-stationary multi-
0
plier; f 0 , f 1 , f 2 is functions that take into account the variability of lubricant
viscosity over the film thickness; λ, σ is relative length and width of the cushion;
ψ = h 20 /(Rav θ ) is relative thickness; θ is the angular length of the cushion with
the IPG; ω, ω∗ is current and characteristic (usually maximum) angular veloci-
ties of the collar; Rav is the average radius of the pad; h 20 is the characteristic
thickness of the bearing film; μ0 is viscosity at the temperature t0 of the lubricant
supply to the IPG; τ = τ/τ∗ is dimensionless time;
2 the internal energy balance equation describing the three-dimensional temperature
distribution in the lubricating and boundary films of the L 2 (−1 ≤ r ≤ 1, 0 ≤
ϕ ≤ θ p , θ p ≤ ϕ ≤ 1, 0 ≤ y ≤ 1) area. In a divergent dimensional nonstationary
form, the energy equation takes the following form
( ) ( )
∂t ∂ρ 1 ∂ ∂ cpρ λoil ∂t
cp ρ +t + (c p ρr Vr t) + Vϕ t − 2 +
∂τ ∂τ r ∂r ∂ϕ r r ∂ϕ
( ) [ ( ) ( ) ] (2)
∂ ∂t ∂ Vϕ 2 ∂ Vr 2
+ c p ρVy t − λoil =μ + ,
∂y ∂y ∂y ∂y
210 N. Sokolov et al.

where t is the local temperature, c p , λoil is the isobaric heat capacity and thermal
conductivity of the lubricant, and ρ is the local density of the lubricant. Equa-
tion (2) was transformed into a dimensionless form at the stage of numerical imple-
mentation using a dimensionless temperature t = c po ρ0 h 220 (t − t0 )/(μ0 ω∗ Rav2
θ)
while maintaining the divergent form;
3 the three-dimensional distribution of temperatures in the areas of the pad L 3 (−1 ≤
r ≤ 1, 0 ≤ ϕ ≤ θ p , 0 ≤ y p ≤ 1) and the rotating collar L 4 (−1 ≤ r ≤ 1,
0 ≤ ϕ ≤ 1, 0 ≤ y c ≤ 1) is described by its own heat transfer equations with
the corresponding boundary conditions: at the outer boundaries, heat transfer is
taken into account by the Newton-Richmann boundary conditions; between the
lubricating film and the pad, as well as between the lubricating and boundary
films and the collar, the conditions for the continuity of temperatures and heat
fluxes (conjugation condition) are set.
The lubrication rates V r and V ϕ are derived from the truncated Navier–Stokes
equations after estimating dimensionless quantities using the N.A. Slezkin method
[15] and taking into account the condition that the pressure gradient along the clear-
ance height h is equal to zero. In contrast to previous studies [13], the velocity V y was
obtained by solving the truncated Navier–Stokes equation along the axis y, which in
a dimensionless form takes the following form
( )
∂V y 2 ∂ ∂V y
Reψ ρ Sh
3
= 2ψ μ −
∂τ ∂y ∂y
[ ( )] (3)
2ψ 2 ∂ λ ∂ [ ] 1 ∂V ϕ ∂V y
− μ (σ r + 1)V r + + .
3 ∂y (σ r + 1) ∂r (σ r + 1) ∂ϕ ∂y

The boundary conditions for Eq. (3) in the area of the lubricating film are assumed
to be equal to zero velocity V y on the surfaces of the pad and collar (impermeability
condition). In the boundary film of the inter-pad groove at θ p ≤ ϕ ≤ 1, y = 1, the
velocity is calculated through the outer normal ν→ to the conditional boundary film
boundary. In dimensionless form, the condition takes the following form
[ ]
∂h 1 ∂h
V y = ψ λ Vr + Vϕ . (4)
∂r σ r + 1 ∂ϕ

When modeling the operation of a hydrodynamic bearing, many authors, due to


mathematical complexity, usually use the calculation of the temperature at the inlet to
the lubricating film at ϕ = 0 by two methods: using the heat balance equation when
setting the temperature difference at the inlet and outlet of the bearing [3], as well
as based on the mixing model in the inter-pad groove in accordance with the energy
balance of three flows and the introduction of the hot oil entrainment coefficient λ,
which varies depending on the bearing design (method of supplying lubricant to the
film) and operating parameters [8, 16, 17]. In the PTEHD mathematical model, the
boundary condition for the energy Eq. (2) at the inlet to the lubricant film is expressed
Numerical Study of the Lubricant Viscosity Grade Influence on Thrust … 211

in a periodic form, which implies the equality of temperatures and heat fluxes of the
lubricant (due to convection and thermal conductivity). In dimensional form, in the
absence of skew of the pads and the beating of the collar, the condition takes the
following form
( )| ( )|
cpρ λoil ∂t || cpρ λoil ∂t ||
t|ϕ=0 = t|ϕ=θ , Vϕ t − 2 = Vϕ t − 2 (5)
r r ∂ϕ |ϕ=0 r r ∂ϕ |ϕ=θ

The differential equations of the mathematical model in the PTEHD formulation


and their boundary conditions are interconnected through such physical properties
of the working lubricant as viscosity, density, heat capacity and thermal conductivity,
as well as through the shape of the clearance, which includes the geometric profile
of the active surface of the pad, and some regime parameters.

3 Construction of a Grid Scheme of the Discontinuous


Galerkin Method

For numerical solution of Eq. (1), a scheme of the summatorial identities method was
constructed, and for Eq. (2), a scheme of the Galerkin discontinuous method with
rectangular elements. The heat transfer equations of the pad and collar are solved
by finite element methods (FEM). In the constructed scheme for Eq. (2), piecewise
constant within the computational domain Ωh and piecewise linear near the boundary
┌ y of the space of approximating functions were used. The choice of this type of
approximating functions allows us to significantly reduce the number of resources
required for the calculation. The way of constructing schemes of this type is given
in [18]. The mesh scheme in the operator form at a fixed ρ form has the following
form
( )
∂(ρh u h )
B + (Av + Aq + Aγ )u h = F + Fγ , (6)
∂τ

where ρh , u h are the grid approximations of the density and temperature functions,
respectively. The operators in Eq. (6) are defined by the following forms
∑∮ ( )
A v u h · wh = −u h V · ∇wh d x+
K∈ h K
∑∮ [ ( )− ( )+ ]( ) (7)
+ u h,+ p V · p − u h,− p V · p wh,+ p − wh,− p d x,
γ \┌ y γ

∑∮ ∑∮ ( )
A q u h · wh = qh · ∇wh d x + wh,+ p − wh,− p qh,+ p · pd x (8)
K∈ h K γ \┌ y γ
212 N. Sokolov et al.

∑∮ ∑∮
A γ u h · wh = ωα u h wh d x, Bu h · wh = buwh d x, (9)
γ ∈┌ y γ K∈ h K

∑∮ ∑∮
F · wh = f wh d x, Fγ · wh = λwh d x. (10)
K∈ h K γ ∈┌ y γ

The operators in Eq. (6) are defined for any ∀wh from the space of approximating
functions, w = (wh,r , wh,ϕ , wh,y ). In equalities (7)-(10) h is a set of partition
elements of the area Ω; K ∈ h is partition element; p is a unit normal to the
boundaries of √the elements of the partition of the area, oriented so that e · p > 0,
e = (1, 1, 1)/ 3; qh is grid approximation of the dissipative part of the equation;
(w)± is the positive or negative part of the w function. It should be noted that the
spaces of approximating functions generally contain discontinuous functions. The
symbols w± p denote the limiting values of the functions of the partition elements
adjacent to the boundary from the side ± p.
To solve problem (6), an iterative method is constructed
( )
ρ k u k+1 − ρuh
B h
+ ( Av + Aq + A g )u k+1
h = F + Fg , k = 0, 1, . . . , (11)

where the symbol w denotes the value of the function w from the previous time layer,
and as u 0h is taken the initial value for the energy Eq. (2) in the lubricating film. To
solve the resulting system of Eqs. (1)–(3), etc., the stabilized method of biconjugate
gradients with incomplete LU decomposition is used as a preconditioner [13].

4 Results of Numerical Experiments

Before carrying out numerical calculations, the following values were taken as initial
data [13]: inner and outer diameters are D1 = 70 mm and D2 = 115 mm; number
of pads is Z = 8; the angular length of the bevel and pad are θbev = 29.1° and θp =
38.8°; bevel width and depth are hbev = 20 mm and δbev = 0.05 mm. The thickness
of the rotating collar is Hc = 25 mm, and the thickness of the pad is Hp = 5 mm.
The lubricant supply temperature is t0 = 40 °C. At the conditional boundary of the
boundary film, the temperature of the lubricant in the IPG is set, i.e. t = t0 . The sizes
of the approximating grids are Nr = 51, Nϕ = 71, Ny = 51, Nyp = 9, Nyc = 9.
As a lubricant in the experiments, the characteristics of turbine oils with different
viscosity grades were used: oil Tp-22S according to the standard TU 38.101821–83
or Tp-22B according to the standard TU 38.401–58-48–92 (ISO VG32), as well as
Tp-30 according to standard GOST 9972–74 (ISO VG46). Some of their previously
measured physical characteristics are listed in Table 1.
Numerical Study of the Lubricant Viscosity Grade Influence on Thrust … 213

Table 1 Physical characteristics


Viscosity Viscosity at Viscosity at Density at Specific heat Viscosity
grade 40 °C, mm2 /s 100 °C, mm2 /s 40 °C, kg/m3 capacity at grade
according to (cSt) (cSt) 40 °C, according to
ISO 3448 ISO 3448
ISO VG32 36 4.4 882.74 1923 0.128
ISO VG46 48.9 5.94 886.2 1906 0.13

The calculations are carried out in dimensional form to facilitate the description of
general patterns. The results of numerical experiments are presented as a dependence
of the maximum temperature tmax of the lubricating film (Fig. 2a) on the circumfer-
ential velocity at different clearance heights (equivalent to the applied axial load) for
oils of two viscosity grades: ISO VG32 and VG46. The circumferential velocity is
determined on the average radius of the pads Rav = 92.5 mm. A parallel taper land
was chosen as the pad profile (Fig. 1a); for a taper land (Fig. 1b), the results are
similar. Thrust bearing power losses are determined by integrating the dissipative
energy dissipation function D over the volume of the lubricating and boundary films,
i.e. the right side of the energy Eq. (2).
As can be seen from Fig. 2a, the temperature tmax increases significantly with
an increase in circumferential velocity. This is due to the release and accumulation
of heat due to the irreversible dissipation of mechanical energy during the viscous
shear of the lubricant layers as it moves in the direction of collar rotation, which
leads to an increase in the temperature level along the lubricant flow [19]. At the
same lubricant supply temperature t0 = 40 °C, oil enters the working area of the
bearing with different viscosities, characterized by a viscosity grade [6–8, 20]. In
this case, the viscosity of the lubricant is a value on which the total heat release due

a) b)

Fig. 2 Dependence of temperatures on the average circumferential velocity of rotation, t0 = 40 °C:


a tmax at different clearances h2 ; b t75/75 at h2 = 50 μm
214 N. Sokolov et al.

to shear in the lubricating and boundary films of the bearing depends [3, 11, 17, 19].
The resulting difference between the temperature curves of oils of viscosity grades
ISO VG32 and VG46 clearly characterizes this pattern. From Fig. 2a, you can also
see that the temperature curve of the lighter ISO VG32 oil looks flatter due to less
heat generation. However, the very difference between the temperature curves of
oils of two viscosity grades, the influence of circumferential velocity and working
clearance has little effect, similar to studies [6, 7]. The difference between the curves
with an increase in the average circumferential velocity gradually becomes constant
and reaches 13° with a clearance of 25 μm, 17° with a clearance of 50 μm, and
11° with a clearance of 90 μm. With a decrease in the height of the clearance h2
between the pads and the collar (see Fig. 1), the maximum temperature increases.
For example, for a thicker ISO VG46 oil with a minimum allowable clearance of
25 μm [3, 17], the temperature tmax already reaches the limit value of 113 °C at a
velocity of 24.22 m/s (corresponding to a rotational speed of 5000 rpm). However, for
ISO VG32 oil, due to reduced heat generation, the bearing operation range is more
extended: with a clearance of 25 μm, the temperature reaches 133 °C at a velocity of
33.9 m/s (corresponding to a rotational speed of 7000 rpm). Thus, the release of heat
in the lubricating and boundary films of the thrust bearing depends on a combination
of three quantities: the rotational speed of the collar (i.e., circumferential velocity),
the operating clearance, temperature, and the viscosity of the supplied oil itself.
In the practice of diagnosing the operation of a thrust bearing, it is recommended to
install a temperature sensor or thermocouple at the 75/75 standard point proposed by
Elwell [19, 21] and prescribed in API 670 (4th edition): the sensor should be located
at a distance of 75% of the width of the pad radially from the inner diameter (i.e. r =
51.875 mm) and 75% of the pad length from the leading edge (i.e. ϕ = 0.508 rad.).
As can be seen from Fig. 2b, with this method, the sensor does not actually measure
the maximum temperature: with an increase in the average circumferential velocity,
the difference between the temperature curves tmax and t75/75 increases. With a
clearance of 50 μm for ISO VG32 oil, the difference reaches 59°, for ISO VG46 oil
−71°. Naturally, this is due to the location of the maximum temperature tmax of the
lubricating film near the active surface of the thrust bearing pad. To understand this,
it is necessary to consider the influence of the determining factors on the temperature
field (isotherms) of the pad.
It is also of interest to consider the change in power losses depending on the
height of the clearance for different viscosity grades. As can be seen from Fig. 3a,
with an increase in the circumferential velocity and a decrease in the clearance h2 ,
the bearing power losses increase significantly, which is associated with an increase
in the velocity gradients Vr and Vϕ in the height of the clearance between the pads
and the collar. For the thicker ISO VG46 oil, power losses are increased, which can
be explained by an increased viscosity index μ of the incoming oil and significant
shear losses. For ISO VG46 oil, at different clearances and at a certain combination
of circumferential velocities, the losses can coincide (point A). It should be noted that
the paper considers only hydrodynamic power losses in the lubricating and boundary
films of the bearing without considering the losses due to turbulence in the IPG and
Numerical Study of the Lubricant Viscosity Grade Influence on Thrust … 215

a) b)

Fig. 3 Dependence of power losses a and bearing capacity b of the thrust bearing on the average
circumferential rotation velocity at different clearances h2 and viscosity grade, t0 = 40 °C

the parasitic friction of the free surfaces of the rotating collar against the moving
lubricant, the contribution of which to the total losses can be significant [3, 6, 7].
The change in the bearing capacity of the thrust bearing is also considered: an
increase in circumferential velocity increases the load capacity only at small clear-
ances (Fig. 3b). For ISO VG32 oil, a further increase, on the contrary, leads to some
reduction in losses. With an increase in the clearance h2 , the influence of circumferen-
tial velocity and oil viscosity grade (i.e. the viscosity of the oil being surrounded at a
specified temperature) becomes minimal. Apparently, this is due to the mutual influ-
ence of several mechanisms. For example, with increased clearances due to reduced
hydraulic resistance, the lubricant flow moves between the pads and the collar at
reduced velocities and viscosity. However, the inevitable release of heat during shear
losses, in turn, leads to a decrease in viscosity (dilution) of the working volume of
the oil and a decrease in heat generation. The shape of the bearing capacity curve
depends on the combination of these inverse mechanisms. With a clearance of 50 μm
at a velocity of more than 28 m/s, there is some difference between the two viscosity
grades up to 10%. With a clearance of 90 μm, the bearing capacity of a thrust bearing
in the two viscosity grades ISO VG32 and VG46 is almost the same.
The following articles will present the results of changing the main characteristics
of the thrust bearing under dynamic loading, i.e. during axial movement of the rotating
collar, as well as in the modes of starting and stopping (running out) of the rotor.
216 N. Sokolov et al.

5 Conclusions

Based on the numerical studies carried out using the Sm2Px3Txτ calculation
program, the following conclusions can be drawn:
1. The maximum temperature tmax of the lubricating film increases significantly with
an increase in the average circumferential rotation velocity due to the release and
accumulation of heat during the dissipation of the lubricant. The release of heat in
the lubricating and boundary films of the thrust bearing depends on the velocity
of the collar (i.e. circumferential velocity), the size of the working clearance,
temperature and the viscosity of the oil supplied directly.
2. The temperature regime of the thrust bearing is significantly affected by the
viscosity grade of the supplied oil (oils of viscosity grades ISO VG32 and VG46
are considered). At the same time, the influence of the velocity itself and the
height h2 of the clearance becomes insignificant on the difference between the
temperature curves of the two grades of oil with an increase in circumferential
velocity.
3. The use of a thicker ISO VG46 oil leads to higher thrust bearing power losses
than ISO VG32 oil: at a circumferential velocity of 29 m/s, the difference reaches
16%. With a further increase in velocity at some clearances, the difference can
be significantly reduced.
4. The bearing capacity of the thrust bearing depends on the circumferential velocity
only at minimum clearances (up to 20–25 μm). With an increase in the clear-
ance between the pads and the collar, the bearing capacity due to internal heat
exchange and hydrodynamic processes becomes practically independent of the
collar velocity and viscosity grade.
5. In general, ISO VG46 viscosity grade oil is recommended for use at low rotor
speeds of a centrifugal or screw compressor, when power losses and overall low
temperature operation are not so great. With increasing circumferential veloci-
ties and energy consumption, an ISO VG32 viscosity grade oil should be used.
However, in this case, it is necessary to more accurately calculate the minimum
thickness of the lubricating film for a given axial load due to the reduced load
capacity (at least 20–25 μm).

References

1. Harnoy, A.: Bearing design in machinery. Marcel Dekker, New York (2003)
2. Capitao, J.W., Gregory, R.S., Whitford, R. P. Effects of high-operating speeds on tilting pad
thrust bearing performance. Journal of lubrication, Tech. Jan. 98(1), pp. 73–79 (1976). https:/
/doi.org/10.1115/1.3452779
3. Khadiev, M.B., Khamidullin, I.V. Compressors in technological processes. Calculation of plain
bearings of centrifugal and screw compressors. Kazan: KNITU, Russia. 260 p. (2021)
4. Klamann, D. Lubricants and related products: Synthesis, properties, applications, international
standards. Verlag Chemie. 489 p. (1984)
Numerical Study of the Lubricant Viscosity Grade Influence on Thrust … 217

5. Wang, Q.J., Chung, Y.-W. Encyclopedia of tribology. Springer Science + Business Media New
York (2013). https://doi.org/10.1007/978-0-387-92897-5
6. Glavatskih, S., Scan, M. Influence of oil viscosity grade on thrust pad bearing operation.
Proceedings of the Institution of Mechanical Engineers, Part J: Journal of Engineering
Tribology, Vol. 5, No. 218, 1994–1996, pp. 401–412 (2004). https://doi.org/10.1243/135065
0042128085
7. Brockwell, K., Dmochowski, W., Scan, M.: An investigation of the steady-state performance
of a pivoted shoe journal bearing with ISO VG 32 and VG 68 oils. Tribol. Trans. 47, 480–488
(2004). https://doi.org/10.1080/05698190490493382
8. Glavatskih, S., Fillon, M., Larsson, R.: The Significance of oil thermal properties on the perfor-
mance of a tilting-pad thrust bearing. J. Tribol. 124(2), 377–385 (2002). https://doi.org/10.1115/
1.1405129
9. Fedotov, P.E., Fedotov, E.M., Sokolov, N.V., and Khadiev, M.B. Sm2Px3Txτ – Dynami-
cally loaded thrust plain bearing when formulating a direct problem. Certificate of the state
registration of a computer program, No. 2020615227, Russia (2020)
10. Khadiev, M.B. Hydrodynamic, thermal and deformation characteristics of lubricating films of
support and sealing units of turbomachines. Dissertation for the Doctor of technical sciences,
Kazan State Technological University, Russia. 410 p. (2002)
11. Hori, Y. Hydrodynamic lubrication. Springer-Verlag, 231 p. (2006)
12. Sokolov, N.V., Khadiev, M.B., Maksimov, T.V., Fedotov, E.M., and Fedotov, P.E. Mathematical
modeling of dynamic processes of lubricating layers thrust bearing turbochargers. Journal of
Physics: Conf. Ser., Vol. 1158, No. 04219 (2019). https://doi.org/10.1088/1742-6596/1158/4/
042019
13. Sokolov, N.V., Khadiev, M.B., Fedotov, P.E., and Fedotov, E.M. Three-dimensional periodic
thermoelastohydrodynamic (PTEHD) modeling of hydrodynamic processes of a thrust Bearing.
International scientific and technical engine conference (EC 2021), IEEE Xplore (2021)
14. Dowson, D.: A generalized Reynolds equation for fluid-film lubrication. Int. J. Mech. Sci. 4(2),
159–170 (1962). https://doi.org/10.1016/S0020-7403(62)80038-1
15. Slezkin, N.A. Dynamics of a viscous incompressible fluid. State publishing house of technical
and theoretical literature, Moscow, Russia. 520 p. (1955)
16. Heshmat, H., Pinkus, O.: Mixing inlet temperature in hydrodynamic bearings. J. Tribol. 108,
231–244 (1986). https://doi.org/10.1115/1.3261168
17. He, M., Byrne, J.M., Cloud, C., and Vazquez, J. Steady state performance predictions of directly
lubricated fluid film journal bearings. Proceedings of the 41st Turbomachinery Symposium.
Texas A&M University. Turbomachinery Laboratories. Huston, Texas (2012). https://doi.org/
10.21423/R1PM0P
18. Fedotov, E.M.: Limit Galerkin-Petrov schemes for the nonlinear convection-diffusion equation.
Springer Link, Differential equations 46(7), 1042–1052 (2010). https://doi.org/10.1134/S00
12266110070116
19. He, M., Byrne, J.M. Fundamentals of fluid film thrust bearing operation and modeling. Asia
Turbomachinery and Pump Symposium, pp. 1–26. (2018)
20. Mang, T. Encyclopedia of lubricants and lubrication. Springer-Verlag, Berlin Heidelberg
(2014). https://doi.org/10.1007/978-3-642-22647-2
21. Elwell R.C. Thrust bearing temperature. Machine design, pp. 79–81 and 91–94 (1971)

You might also like