Você está na página 1de 14

Tribology Transactions

ISSN: 1040-2004 (Print) 1547-397X (Online) Journal homepage: http://www.tandfonline.com/loi/utrb20

Numerical Simulations of Lubricating Grease


Flow in a Rectangular Channel with and without
Restrictions

Chiranjit Sarkar, Lars G. Westerberg, Erik Höglund & T. Staffan Lundström

To cite this article: Chiranjit Sarkar, Lars G. Westerberg, Erik Höglund & T. Staffan Lundström
(2017): Numerical Simulations of Lubricating Grease Flow in a Rectangular Channel with and
without Restrictions, Tribology Transactions, DOI: 10.1080/10402004.2017.1285090

To link to this article: http://dx.doi.org/10.1080/10402004.2017.1285090

Accepted author version posted online: 20


Jan 2017.
Published online: 20 Jan 2017.

Submit your article to this journal

Article views: 32

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=utrb20

Download by: [Indian Institute of Technology - Patna] Date: 02 March 2017, At: 23:36
TRIBOLOGY TRANSACTIONS
http://dx.doi.org/10.1080/10402004.2017.1285090

Numerical Simulations of Lubricating Grease Flow in a Rectangular Channel with


and without Restrictions
€glundb, and T. Staffan Lundstro
Chiranjit Sarkara, Lars G. Westerberga, Erik Ho € ma
a   
Division of Fluid and Experimental Mechanics, Lulea University of Technology, Lulea, Sweden; bDivision of Machine Elements, Lulea University of

Technology, Lulea, Sweden

ABSTRACT ARTICLE HISTORY


This article presents numerical simulations of the laminar flow of lubricating greases in a channel with Received 12 September 2016
rectangular cross section. Three greases with different consistencies (NLGI grades 00, 1, and 2) have been Accepted 16 January 2017
considered in three different configurations composed of a rectangular channel without restrictions, one KEYWORDS
rectangular step restriction, and one double-lip restriction. The driving pressure drop over the channel Lubricating grease flow;
spans from 30 to 250 kPa. The grease rheology is described by the Herschel-Bulkley rheology model, and boundary layer; velocity
both the numerical code and rheology model have been validated with analytical solutions and flow profile; rheology; lubrication;
measurements using micro-particle image velocimetry. computational fluid
dynamics (CFD)

Introduction
with double restricting sealings in focus (Green, et al. (9); Baart,
Lubricating greases are used to lubricate various machine ele- et al. (10)). Direct measurements have also been done by Britton
ments such as rolling bearings, seals, and gears. Compared to oil, and Callaghan (11) and Magnin and Piau (12), who visualized the
grease has many advantages due to its consistency because it grease flow in a plate–plate viscometer using nuclear magnetic res-
adheres to the solid surfaces of the geometry preventing leaking onance microscopy, confirming a uniform shear flow across the
and corrosion. Understanding the flow dynamics of grease is gap for Newtonian fluids and an anomalous behavior for non-
crucial for optimum design of an arbitrary confinement geome- Newtonian fluids. The shear rate was averaged across the gap, but
try and for the understanding of lubrication mechanisms. How- this was valid only at relatively high shear rates and not for low
ever, grease flow in a rolling element bearing, for example, is shear rates. At low shear rates, a plug flow in combination with pos-
very complex due to large variations in flow conditions in vari- sible wall slip occurred, which gave rise to a variation of shear rate
ous parts of the bearing, with both wall bounded and free-surface across the gap. To measure this in detail, Westerberg, et al. (1) stud-
flows present. Hence, the rate of shear varies enormously with a ied the slip phenomena at low shear rates using mPIV. They mea-
span of around eight decades depending on the location within sured the velocity profiles and slip was indicated by large velocity
the bearing. Considering the motion of the grease from the gradients in a thin layer adjacent to the wall. An analytical model
beginning of the bearing operation, churning takes place, and based on the Herschel-Bulkley rheology model was also developed
through the motion of the rolling elements, the grease is pushed for validation.
to the sides of the bearing, onto the seals/shields, and onto the Having validated numerical models of the flow in a bearing
bearing cage. In the vicinity of the rolling elements, the grease during operation, including start and stop and phase separa-
will act as a reservoir, enabling relubrication of the bearing and tion, would be a great tool in the design of bearings and for the
it is the flow motion of grease in the bearing, which determines understanding of the lubrication mechanism. This is, however,
the reservoir formation (Westerberg, et al. (1)). The non-Newto- far away, and numerical models of lubricating grease flow are
nian nature of grease rheology (Barnes (2)) increases the com- scarce in the scientific and technical literature. This is not the
plexity of the flow behavior. In addition, phase separation case for oil lubrication; the main reason is that oil is a single-
causing wall slip (Czarny and Moes (3); Czarny (4); Bramhall phase fluid with a Newtonian rheology behavior. Lubricating
and Hutton (5); Vinogradov, et al. (6)) adds to the complexity. greases have a non-Newtonian, shear-thinning rheology due to
Measuring the grease flow in a real bearing is very difficult and their two-phase composition composed of a base oil and thick-
has not been performed so far. To simplify the problem, qualitative ener agent. Lithium hydroxy stearate is the most widely used
measurements in rectangular channels have been done by various thickener (Lugt (13)). In addition, the phases separate depend-
researchers (Westerberg, et al. (1); Li, et al. (7); Scarlett (8)) and ing on the flow conditions and possibly other effects, implying
direct flow measurements have been performed on grease in rotat- gradients of oil and thickener in the flow, which in turn induce
ing geometry with microparticle image velocimetry (mPIV) but mechanisms such as wall slip and shear banding.

CONTACT Chiranjit Sarkar chiranjit.sarkar@ltu.se


Color versions of one or more of the figures in the article can be found online at www.tandfonline.com/utrb.
Review led by Michael Dube.
© 2017 Society of Tribologists and Lubrication Engineers
2 C. SARKAR ET AL.

Lubricant flow in a journal bearing supply pocket using a and 150 kPa, respectively, for the three greases (NLGI00,
PIV technique has been performed by Kosasih, et al. (14). They NLGI1, and NLGI2) to match the study by Westerberg, et al.
demonstrated the use of PIV on difficult geometrical configura- (1). In this study, a physics-controlled mesh with an extra fine
tion and proposed a correction procedure when an endoscope mesh size was used to discretize the fluid domain. The total
is used. They also aimed to shed some knowledge about flow number of elements is 300,904 considering 285,064 triangular
field in journal bearing supply pocket. However, the flow field elements and 15,844 quadrilateral elements. There are 8,056
inside the journal bearing is missing. The present article aims edge elements and four vertex elements. The pressures at the
at being a first step toward numerical models of lubricating outlet, pout, is kept as 0 Pa. A steady 2D, single-phase laminar
grease flow during running conditions in a rolling element flow module is used with a PARIDOS scheme to solve the sys-
bearing configuration. Here a single-phase numerical grease tems of equations. Comsol has the power law, Carreau, and
model is built and validated with flow measurements in a rect- Cross rheology models (15) implemented. The Carreau model
angular channel without restrictions (Westerberg, et al. (1)) reads
and corresponding measurements of the velocity profiles in a
rectangular channel with a step and double-lip restriction (Li, h0 ¡ h 1
h D h1 C n¡1 ; (3)
et al. (7)). Using the numerically developed models, particle ½1 C ðK:g_ Þ2  2

flow in a restricted channel is investigated.


where h 1 is the viscosity plateau at high shear rates, h0 is the
Methods viscosity plateau at low shear rates, g_ is the shear rate, K is the
grease consistency, and n is the power law exponent. Consider-
Governing equations, rheology model(s), and greases ing the viscosity plateau at high shear rates, h 1 as zero, Eq. [3]
Comsol Multiphysics v5.2 (15) has been used for the numerical yields
modeling. Comsol enables user-defined expressions of the appar- h0
ent viscosity as a function of the built-in shear rate (denoted spf. hd D ; (4)
sr), meaning that any rheology model can be implemented. As 1 C ðK:g_ Þ1 ¡ n
introduced earlier, the main objective of the numerical modeling
is to validate the model with measured velocity profiles and which is the Cross rheology model. In this article, however, the
developed analytical expressions, thereby ensuring the quality modified Herschel-Bulkley (H-B) rheology model (Misoulw
and validity of the numerical simulations including the accuracy and Abdali (16)) is applied. It reads
of the applied rheology model. With a working model, the analy- t0 
sis can be applied to more complex geometries and can also be hD 1 ¡ e ¡ m:g_ C K g_ n ¡ 1 ; (5)
g_
extended to flow cases not covered in the experimental work.
The grease is described as a single-phase continuous fluid.
where t0 is the yield stress, g_ ( D grad(u) C (grad(u)T) the
The equations governing the flow are the equation of motion
shear rate, K is the grease consistency, n is the power law expo-
(the Navier-Stokes equations) and the continuity equation
nent, and m is a constant.
respectively. For a stationary flow these reads (here shown as
The H-B rheology model is shown to work well except for at
implemented in Comsol)
the zero and infinite shear rate plateaus (see, e.g., Lugt (13), p.
  112) and, based on this the model, fits well for the actual flow
rðurÞu D r½ ¡ pI C m ru C ðruÞT  C F cases in the present study. For flows at really high shear rates
(1a--1b)
ru D 0; the Carreau model would be a better choice because it also con-
tains the base oil viscosity, which dominates the grease rheol-
where u is the velocity vector, r is the density, m is the dynamic ogy at high shear rates. The Carreau and Cross models also
viscosity, p is the pressure, I is the identity tensor, and F is the contain the zero shear rate viscosity, making them suitable for
volume force(s)—typically gravity and/or centrifugal forces. creep flow.
The continuity equation (Eq. [2]) yields that the divergence An important issue in numerical modeling in general is
term in Eq. [1a] vanishes. Comsol enables user-defined rheol- quality and trust. The unrestricted flow case has been modeled
ogy models, which in turn introduces a great flexibility in the analytically and validated with experimental data. The H-B rhe-
modeling work. The boundary conditions at the inlet and outlet ology model is well suited for analytical modeling because it is
are set to fixed pressures with no viscous stresses; that is, p D fairly straightforward and has consequently been considered
pin and p D pout at the inlet and outlet, respectively. In addition, for the analytical modeling work. Using the H-B model for the
no viscous stresses in the normal direction are applied, mean- numerical modeling work, we hence have great sources for vali-
ing dation enabling reliable numerical models of more complex
flows.
  The modified H-B model describes the same relation
n½m ru C ðruÞT  D 0: (2) between the shear rate and the shear stress as the correspond-
ing basic form. However, in the modified version of the rheol-
These boundary condition equations hold at the channel ogy model the factor in brackets is added. The H-B model is
wall. In order to study the effect of the pressure drop on the overall a good engineering model valid in the shear thinning
flow a parametric solver has been used to vary pin as 50, 100, regime considering the shear rate–viscosity relationship (Lugt
TRIBOLOGY TRANSACTIONS 3

(13)) but it is physically rather crude because the flow in the Figure 1 shows the rectangular channel with length L D 40 mm
model is considered nonyielded (not moving) if the shear stress in the x-direction, width w D 1.5 mm in the y-direction, and
in the material is less than the threshold yield stress value and height h D 1 mm in the z-direction. The full 3D geometry is
continuously sheared (fully yielded) for shear stress values shown in Fig. 1a and a corresponding 2D view is shown in
above the yield stress. This singular, discontinuous, transition Fig. 1b. In Westerberg, et al. (1) an analytical model of the flow
description has some obvious physical constraints that will lead has been developed; there, however, a 1D flow is assumed; that
to problems if applied numerically. In order to implement the is, the influence of the side walls is neglected as the distance
transition from unyielded to yielded grease, the modified H-B between the top and bottom channel walls in the y-direction is
model has consequently been considered. Here the transition is considered much smaller than the distance between the side
applied through the term in brackets. The constant m sets the walls in the z-direction. This means that @/@y >> @/@z, which
sharpness of the transition. Through a match between the in turn enables a flow model where the velocity only is depen-
numerical results with flow measurements, a value of m D 1 is dent on the y coordinate; that is, a 1D flow. In Westerberg,
used. et al. (1), the 1D flow model is motivated by the plug flow char-
In this study, the numerical models are validated against acter of the grease flow. Because the cross section of the channel
flow measurements of three lithium greases with NLGI grades is close to square it is of specific interest to investigate the influ-
00, 1, and 2, respectively. The rheological data based on the ence of the geometry on the flow compared to the ideal 1D
Herschel-Bulkley rheology model are presented in Table 1. case. This is done in the present study by numerically modeling
Here our focus is on the relation between grease rheology and the ideal 1D case and comparing the velocity profiles with the
grease flow, not the greases as such. The impact of grease com- measured velocity profiles and the analytical model in Wester-
position on the flow is indeed highly relevant and interesting berg, et al. (1).
but is left for future work. Three typical greases used in bear- In addition to comparing the influence of the neglected side
ings are consequently used in the present study. walls in the 1D analytical model, the numerical model enables
investigation of the location of the focal plane where the flow
measurements are made. One difference between mPIV and
Flow in rectangular channel: Validation with analytical macro PIV is that in the former the measurement plane is set
model and data from flow measurements using mPIV by the focal depth of the microscope as the whole geometry
Microparticle image velocimetry (channel) is illuminated by the laser light. In PIV, the measure-
The PIV technique is a nonintrusive method to capture the ment plane is set by imaging optics forming a thin laser sheet
motion of a flowing medium. The main principle is that images analogous to the focal plane in the microscope. In the work in
of a particle-seeded flow are taken by a high-speed camera. a Westerberg, et al. (1), the location of the focal plane has not
pulsed laser is used as the light source. The laser is synchronized been determined. In order to determine the location of the focal
with the camera in order to capture images separated by a small
time step. The tracer particles are illuminated in a single plane
of the flow with a light sheet whose thickness is less than the
depth of field of the image recording system. In mPIV, the plane
of measurement is set by the focal depth of the microscope. This
means that the velocity profiles are measured in this plane; 3D
PIV systems measuring the velocity component out of the plane
also exist. For the measurements referred to in this article, par-
ticles with a diameter of 7.68 § 0.19 mm have been used. The
particle motion, and ultimately the flow motion, is traced using
a correlation algorithm. For more details about mPIV, the reader
is referred to Westerberg, et al. (1) and Barnes (2).

Rectangular channel without restrictions


The numerical simulations are performed on the same geome-
tries used in Westerberg, et al. (1), where flow measurements of
the flow in a rectangular channel without restrictions are pre-
sented, and in Li, et al. (7), where corresponding measurements
are performed in a rectangular channel with restrictions.

Table 1. Rheological parameters for the grease based on the Herschel-Bulkley rhe-
ological model (Li, et al. (7)).
Grease t 0 (Pa) K (Pa.s) n r (kg/m3) Figure 1. (a) Three-dimensional straight channel used in the mPIV measurements
by Westerberg, et al. (1). (b) View of a 2D straight channel where the side walls
NLGI 00 0 1.85 1 890 not are affecting the flow. The zoomed-in view shows a schematic representation
NLGI 1 189 4.1 0.797 910 of the situation close to the wall with slip present in flow. In the slip layer, the
NLGI 2 650 20.6 0.605 930 velocity gradient is much larger compared to that in the bulk flow (Westerberg,
et al. (1)).
4 C. SARKAR ET AL.

plane used in the flow measurements, the velocity profile from the only velocity component is the one in the main flow direc-
the numerical model at a number of different planes in the tion and only is dependent on the coordinate perpendicular to
geometry is compared to the measured velocity profile from the main flow direction—the distance between the top and bot-
the mPIV measurements. The channel cross section with corre- tom walls (Figs. 1–3) is much smaller than the distance between
sponding planes is shown in Fig. 2. the side walls as described in the previous section. Westerberg,
et al. (1) developed an analytical model for a 1D flow where the
Rectangular channel with restrictions expression for the velocity in the main flow direction (cf.
For validation in a more complex geometry inducing a larger span Fig. 1b) as a function of the coordinate perpendicular to the
of shear rates, two additional channels with step and double-lip main flow direction reads
restriction are modeled (see Fig. 3). The dimensions of the channels
 n C 1
are identical to the previous case apart from the restrictions being n 1 1 dp  w t 0 n
centered at the middle of the channel in the x-direction. For these u.y/ D y ¡ ¡
n C 1 1 dp K dx 2 K
channels, experimental flow measurements in Li, et al. (7) enable K dx
(7)
 n C 1
comparison with velocity profiles at different locations in the n 1 1 dp  w t 0 n
geometry. The pressure drop in the simulations was varied from 30 ¡ ys ¡ ¡ C us :
n C 1 1 dp K dx 2 K
to 250 kPa to match the study in Li, et al. (7). K dx

Here n, K, and t0 are the rheological parameters in the H-B


Results and discussion
rheology model; w is the width of the channel, and dp/dx is the
Westerberg, et al. (1) clarify that it is possible to obtain a good pressure gradient. ys is the thickness of the slip layer (if pres-
match between measured velocity profiles and a developed ana- ent), and us is the corresponding slip velocity. For the case of
lytical model for a 1D flow. A requirement is, however, that the no slip, these values are equal to zero. The expression for the
value of the pressure drop is adjusted together with the rheolog- velocity profile is valid in the region ys yyl .yl w 6 2/, where
ical parameters. One reason for this adjustment, as discussed in yl is the location of the yield point where the shear stress in the
their paper, might be that the rheological measurements are flow is below the grease yield stress; that is, the location where
wrong due to slip in the flow and/or that the measured pressure the plug region starts. The velocity profile in the channel is
drop is incorrect. Another possibility, as addressed in the pres- symmetrical as to y D w/2.
ent article, is that the 1D flow model applied is too crude for The developed analytical expression for the velocity pro-
the experimental flow in the channel with a nearly square cross file is mathematically an exact solution of the flow. This
section. yields that with the same rheology model used in the analy-
sis and in the numerical modeling, the analytical model rep-
resents the truth upon which the numerical model is
Channel without restrictions
validated against. The rheology data presented in Table 1
Validation of numerical grease flow model for the ideal 1D have been considered. The pressure gradient in the analyti-
flow case cal model is calculated from the pressure drops defined in
In order to investigate the impact of the side walls with the numerical model, which in turn are the same as in the
(inward) surface normal in the § z-direction in Fig. 2, an ideal experimental work in Westerberg, et al. (1). The inlet pres-
1D flow is first considered. In an ideal 1D flow—that is, where sures are 50, 100, and 150 kPa for the NLGI 00, NLGI 1,
and NLGI 2 greases, respectively. This corresponds to a
pressure gradient of ¡1.25, ¡2.5, and ¡3.75 MPam¡1,
respectively. Similarly, the slip velocities found in the exper-
imental work in Westerberg, et al. (1) are 0.0073, 0.0021,
and 0.0032 m/s for NLGI 00, NLGI 1, and NLGI 2 greases
respectively. These values of slip velocities are used to plot
Fig. 4. Because the geometry is limited to 2D (x, y), it is
compared to the 1D analytical model. Figure 5 shows the
convergence plot of error with iterations. It confirms that
the level of mesh refinement is sufficient to obtain the con-
vergence plot. Figure 4 shows a comparison between the
analytical and numerical velocity profiles across the channel
for the three greases. The solid line represents the analytical
model and the dotted line is the velocity profile from the
numerical model. As shown in Fig. 4, the match between
the analytical and numerical velocity profiles is excellent.
The numerical model captures the yield behavior of the
grease as shown by the evolution of the plug region when
Figure 2. Cross section view of the rectangular channel used in Westerberg, et al. the NLGI grade increases (see Fig. 4). From these results, it
(1) with the different planes considered for determining the measurement plane
in the flow measurements. A–A: Mid plane; B–B: 0.15 mm from the mid plane; C– is concluded that the numerical model works well for the
C: 0.25 mm from the mid plane; and D–D: 0.40 mm from the mid-plane. 1D flow case.
TRIBOLOGY TRANSACTIONS 5

Figure 3. Rectangular channel with (a) step restriction and (b) double-lip restriction. The red zones in (a) represents regions in which the flow is mainly analyzed. The red
lines in (b) are locations for velocity profiles.

2D flow in a channel without restriction The numerical velocity profile surface plot for the 2D flow is
Having validated the numerical model for 1D flow, the 2D full shown in Fig. 6. The driving inlet pressures are 50, 100, and
flow in the channel without restriction is considered. The 150 kPa for the NLGI 00, NLGI 1, and NLGI 2 greases, respec-
impact of the side walls on the flow is investigated and the tively. The evolution of the plug region with increasing grease
numerical model makes it possible to locate the position of the consistency is clearly shown. Considering the grease shear thin-
focal plane where the experimental flow measurements were ning property, this evolution is analogous for a given grease with
made. The latter is of interest, because the exact location in increasing shear rates. Westerberg, et al. (1) show this in the
Westerberg, et al. (1) is not known. For clarification, a 2D flow measurements with mPIV for increasing pressure gradients
flow—analogous to the 1D flow—means that the flow velocity and in the analytical model for the 1D flow where the width of
component(s) is a function of both coordinates (y and z) per- the plug region is reduced with increasing pressure gradient.
pendicular to the main flow direction. The channel geometry Figure 7 shows the velocity profiles of the NLGI 00 grease at
is, however, 2D and 3D for the 1D and 2D flows, respectively. the different planes in Fig. 2. The numerical results are

Figure 4. Velocity profiles of (a) NLGI 00 at 50 kPa, (b) NLGI 1 at 100 kPa, and (c) NLGI 2 at 150 kPa. Solid: using 1D analytical model, ~: using 1D numerical model.
6 C. SARKAR ET AL.

profiles for the NLGI 1 grease as for the NLGI 00 grease in


Fig. 7. It was found that the side walls have a significant impact
on the flow as for the NLGI00 grease, with a significantly lower
maximum speed, again around five times. The plane that
matches the measured profile is in terms of maximum speed is
also the mid-plane.
When performing the mPIV measurements, the pressure
was measured with a manometer as presented in Westerberg,
et al. (1). The pressure is, however, measured upstream of an
elbow fitting attached to the setup geometry that directs the
flow to the channel inlet. The exact pressure at the channel inlet
is hence not known; pressure losses are likely to occur between
the point of measurement and the channel inlet, meaning that
the pressure at the channel inlet is lower than stated in the sim-
ulations. This in turn yields a reduced velocity of the flow. In
Figure 5. Convergence plot of relative error with number of iterations. Fig. 9, the velocity profile at the mid-plane has been fitted to
the experimental profile by varying the inlet pressure. It was
found that an inlet pressure of 44 kPa makes a good fit with the
compared to the corresponding experimental velocity profile. experiments, meaning that a pressure loss of 56 kPa is present
Figure 7 also compares the velocity profile for the ideal 1D between the point of measurement and the actual channel inlet.
flow. It follows that the side walls of the channels definitely The corresponding analysis for the NLGI 2 grease (Fig. 10)
affect the velocity profile, with a maximum velocity in the flow reveals the same behavior as for the NLGI 1 grease; the shape
that is about five times lower than for the 1D case. The effect is of the velocity profile indicates that the flow measurements
similar for the NLGI 1 and NLGI 2 greases. The difference have been made at the mid-plane for an inlet pressure of
between the planes A–A through C–C is, however, not espe- 150 kPa. The agreement looks quite good for the NLGI 2
cially large for the NLGI 00 grease; the trend is a velocity grease; therefore, there is no need for pressure adjustment.
decrease when going from the mid-plane toward the side wall. However, investigating the suggested adjustment in pressure
This is logical and in line with the surface plots in Fig. 6 and drop for the NLGI 1 grease at 56 kPa, we note that the pressure
the boundary conditions. It can be concluded that the velocity drop for a Herschel-Bulkley fluid for the flow in a pipe of length
profile from plane A–A matches best with the measured veloc- L and diameter D reads Dp D 4Lty/aD (Lugt (13), p. 140). Here
ity profile; that is, the focal plane is likely to be located close to ty is the yield stress and a is the relative radius of the plug in
the mid-plane rather than closer to the side wall (Fig. 2). the pipe flow. Considering the pressure loss for the full channel
The same procedure and analysis is made for the NLGI 1 length ( D 50 mm), an equivalent diameter of 1.3 mm (calcu-
and NLGI 2 greases. Figure 8 shows the corresponding velocity lated from the channel cross section dimensions), and a yield

Figure 6. Surface plot velocity of the three greases: (a) NLGI 00 at 50 kPa, (b) NLGI 1 at 100 kPa, and (c) NLGI 2 at 150 kPa. The color bar shows the velocity magnitude.
TRIBOLOGY TRANSACTIONS 7

Figure 10. Velocity profiles for the NLGI 2 grease at 150 kPa. Dashed (large): From
Figure 7. Velocity profiles of the NLGI00 grease at 50 kPa. The arrow points in the flow measurements. Scale on the left y-axis. Dash-dotted: The A–A plane (Fig. 2),
direction of increasing distance from the mid-plane (plane A–A to C–C). Solid: The numerical results. Scale on the right y-axis for this curve and following ones. Dotted:
ideal 1D case, numerical result. Dashed: Flow measurements (1); Dash-dotted: The B–B plane, Dash-double dotted: The C–C plane, Dashed (small): The D–D plane.
Numerical result of the 2D flow. Velocity profile at the A–A plane (see Fig. 2); Dot-
ted: Same as previous but at the B–B plane; Dash-double dotted: The C–C plane.
the same order as the pressure drop through the whole channel.
It is consequently not unlikely that the obtained values of the
rheological parameters are incorrect as suggested in Wester-
berg, et al. (1).

Figure 8. Velocity profiles for the NLGI 1 grease at 100 kPa. Dashed (large): From
flow measurements. Scale on the right y-axis. Solid: The ideal 1D flow, numerical
results. Scale on the left y-axis for this curve and following ones. Dash-dotted: The
A–A plane (Fig. 2), numerical results. Dotted: The B–B plane, Dash-double dotted:
The C–C plane, Dashed (small): The D–D plane.

stress value of 189 Pa (for the NLGI1 grease), the pressure drop
is of the order of 40 kPa considering an alpha value of 0.75. A
pressure drop between the manometer and the channel inlet of
56 kPa is thus extremely large because the adjusted value is on

Figure 9. Velocity profile of NLGI 1. Solid: From flow measurements at 100 kPa; Figure 11. Velocity field of the NLGI 2 grease at 250 kPa inlet pressure. (a) Before
Dotted: when inlet pressure 43.5 kPa, numerical results; Dashed: when inlet pres- restriction at zone 1. (b) Steady flow in the restriction at zone 2. (c) After restriction
sure 44 kPa; Dot-dashed: when inlet pressure 44.5kPa. at zone 3. The color bar shows the velocity magnitude in meters per second.
8 C. SARKAR ET AL.

Rectangular channel with restrictions Figure 3a shows the locations at three zones upstream of the
step, at the top of the step, and downstream of the step. A no-
Having validated the numerical model for the grease flow in the
slip boundary condition is applied in the numerical model
straight, rectangular channel without restrictions, the flow in a
because slip velocity was not considered in the work by Li, et al.
rectangular channel with restrictions is now considered. As for
(7). Figure 11 shows the velocity field for the NLGI 2 grease at
the previous case, the numerical model is validated against flow
250 kPa (maximum velocity 0.45 m/s). The flow rate in the
measurements with mPIV. Here the work done by Li, et al. (7)
channel is constant, which together with continuity in the flow
is considered. They performed experiments on a rectangular
yields an increased velocity as the step is approached and then
channel with step and double-lip restrictions. All of simulations
a pure straight channel flow. Figures 11a and 11c indicate sym-
are made on the actual 3D channel geometry, identical to the
metry in the flow before and after the step, and in order to
one used in the experiments. The velocity profiles are evaluated
investigate this further the velocity profiles upstream and
at the mid-plane (A–A; see Fig. 3 and previous subsection).
downstream of the flow are plotted as shown in Figs. 12–13.
From Figs. 13a and 13e) it follows that the flow is fully devel-
Validation of numerical grease flow model in channel with a oped upstream of the step and that the flow has retained its
step restriction upstream profile at the corresponding distance downstream of
For validation, the velocity profiles at different locations are the step. Figure 13b shows the transition flow as the upstream
compared to corresponding measured velocity profiles. corner of the step is rounded, and Fig. 13d shows the mirrored

Figure 12. Positions of cut lines (denoted L) for the velocity profile plots. (a) L1–4 upstream of the step. (b) Zone 1: L5–9 before the restriction. (c) Zone 2: L10–13 at the
top of the restriction.(d) Zone 3: L14–18 after the restriction. (e) L19–22 downstream of the step.
TRIBOLOGY TRANSACTIONS 9

Figure 13. Velocity profiles from the numerical model for the NLGI 2 grease (for an inlet pressure of 250 kPa) at the position of the cut lines presented in Fig. 12. Each sub-
figure corresponds to the same set of cut lines as in the corresponding subfigure in Fig. 12.

case when the flow is leaving the restriction. Figure 13c shows for the inlet pressure specified in Li, et al. (7). Following the dis-
the unrestricted flow at the top of the step. These results show cussion in the previous section, the reason behind this result is
that the flow is completely symmetric along a symmetry line a pressure drop between the location where the pressure is
drawn at the centre of the step restriction. The effect of measured and the inlet of the channel. To predict the actual
increased shear on the flow is also shown in the transition as inlet pressure, the model applied inlet pressure has been varied
the flow rounds the corner of the step, going from a fully devel- in the simulation. It was found that a pressure of 175 kPa
oped plug profile, approaching a parabola, and eventually stabi- results in a velocity profile that matches well with the measured
lizing at the top of the step reaching a plug formation once velocity profile.
again. The behavior of thinner grease was found to be analo- In the work by Li, et al. (7), a circulation area close to the
gous in terms of symmetry and development of the velocity upstream inside corner of the step restrictions was observed.
profiles considering the difference in the channel without Investigating this numerically, the flow streamlines have been
restrictions. With a less pronounced yield stress, the plug flow plotted for the three greases and an inlet pressure of 250 kPa;
effect is decreasing with NLGI grade. see Figs. 15a–15c. In Fig. 15d the streamlines are plotted for the
To compare the velocity profile between the computational NLGI 2 grease at a reduced pressure of 30 kPa. Figures 15a–15c
fluid dynamics (CFD) simulation and mPIV measurements, the show a trend where the circulation area decreases with increas-
velocities along L9 (at zone 1) and L15 (at zone 3) have been ing NLGI grade. There is a visible difference between the NLGI
exported. Figure 14 shows a comparison between them; it fol- 00 grease and the NLGI 1 grease in terms of the size of the cir-
lows that there is a discrepancy between the measured velocity culations area, though the difference is negligible between
profile and the corresponding profile from the numerical model NLGI grades 1 and 2. From Fig. 15d it follows that the
10 C. SARKAR ET AL.

restriction and the double-lip restriction in the following sec-


tion are used as simple model geometries to investigate this
type of flow. A major issue in seals is motion of contaminant
and wear particles; circulation zones like the ones presented
here are of interest because particles will be trapped in such
regions. Recalling the shear-thinning property of lubricating
greases, the viscosity is reduced/increases with an increased/
reduced shear, which in this case is induced by the applied
pressure. The nonlinear relation between shear rate and viscos-
ity hence yields a nonlinear development of the Reynolds num-
ber in the flow. In terms of particle transport for the present
geometry, this means that a flow where the grease is creeping
close to the solid boundaries will transport the particles present
in the flow domain.

Grease flow past a double-lip restriction


Continuing with the flow past a double-lip restriction, we recall
Fig. 3b where the distance between the two restriction tips is
1.5 mm, the minimum gap height is 0.9 mm, and the maximum
gap height between the restrictions is 1.5 mm. The maximum
velocity is observed above the two restrictions where the gap
height is small as found in Li, et al. (7). Figures 16a and 16b
show the velocity magnitude at 30 and 250 kPa, respectively,
for NLGI2 grease. In line with the previous results, a higher
Figure 14. Comparison of velocity profiles of the NLGI2 grease between the CFD pressure drop results in a higher velocity of the flow. Of special
simulations and mPIV measurements along (a) L9 and (b) L15 cutlines. ^: Numeri- interest here is to investigate the flow depth because particles
cal results with inlet pressure of 250 kPa, &: mPIV measurements from Li, et al. stuck between the restrictions—depending on the flow depth—
(7), D: numerical results with inlet pressure of 175 kPa.
will either remain stuck or be dragged along with the flow. The
circulation area is larger because the driving pressure has flow depth is defined as the vertical distance the grease flow
reduced for the NLGI 2 grease. between the two restrictions. In terms of application where this
As introduced earlier in this article, transverse flow is impor- type of flow is present, the flow depth is important for how con-
tant in bearings and seals during relubrication. The step tamination particles (like road dust) and wear particles are

Figure 15. Close-up view of the grease flow streamlines in the corner area at 250 kPa inlet pressure. (a) NLGI 00, (b) NLGI 1, (c) NLGI 2, and (d) NLGI 2 at 30 kPa inlet pressure.
TRIBOLOGY TRANSACTIONS 11

Figure 16. Velocity field for the NLGI 2 grease at (pressure drop) (a) 30 kPa, red line shows streamline flow, white color particles position at 300 s; (b) 250 kPa in 0.3 s; (c)
250 kPa in 0.3 s with same range of color bar of as in (a); (d) NLGI 00 grease at 250 kPa in 0.3 s; and (e) NLGI 1 grease at 250 kPa in 0.3 s.

transported. To capture the behavior of grease flow depth at number of particles has initially been packed across the inlet
high pressure, Fig. 16b has been redrawn considering the same cross section (t D 0). For the high-pressure case (250 kPa) the
scale as for the lower pressure (Fig. 16a). It follows that most of flow has been run for 0.3 s, whereas for the low-pressure case
the domain consists of fully yielded grease. Similarly, Figs. 16d (30 kPa) the run time is 300 s, because the particles took less
and 16e show the flow depth of NLGI 00 and NLGI 1 greases, time to pass through the lips for the high pressure. Comparing
respectively. From all surface velocity plots, it can be concluded Figs. 16a and 16b it follows that the particles follow the bulk
that flow depth increases as grease consistency decreases. flow for the lower pressure (equaling low flow rate), whereas
for the higher pressure particles are trapped in the bottom
Particle motion upstream corner of the lips. This behavior is directly coupled to
Considering the application of seals (especially double restric- the velocity distribution in the channel and around the lips and
tion seals) of which the present double restriction geometry is a inertia effects. For the low-pressure case the velocity is very low
model problem, the motion of particles in the grease flow is of in connection to the lips (blue area in Fig. 16a), whereas for a
specific interest by means of the transport of contamination higher pressure the low-velocity area is very low in regions sur-
particles through the seal. In order to investigate particle trans- rounding the lips. These observations yield that an optimum
port through the double-lip geometry in this article, particles particle transport is supported by low flow rates (low pressure
have been introduced to the flow. This is done by applying a drop). Consequently, in order to trap particles in the flow, the
specified number of particles, with specified size, to be distrib- restrictions should be designed such that high-shear zones are
uted through the cross section of the channel inlet. Here par- present.
ticles with 7.68 mm diameter and mass of 7 mg have been Figure 17 shows the velocity profiles for NLGI 00 and NLGI
considered. A scale factor of 5 has been applied in the postpro- 2 greases at position 1 and position 2 as defined in Fig. 3b. It
cessing to more clearly visualize the motion. The maximum can be observed that NLGI 00 grease has a more parabolic flow
12 C. SARKAR ET AL.

Figure 17. Velocity profile measured at position 1 and position 2 in Fig. 3b: (a) NLGI 00 grease, (b) comparison between simulation and mPIV of NLGI 00 grease, (c) NLGI 2
grease, and (d) comparison between simulation and mPIV of NLGI 2 grease. The flow rate q is the flow across the channel height for each corresponding position.

compared to the NLGI 2 grease through position 1, as suspected dimensional model does not accurately describe the
based on previous results. There is a close match between simu- velocity profile in a channel with rectangular cross sec-
lated and mPIV velocity profiles at the same flow rate. tion of 1.5 £ 1 mm.
3. The modified H-B rheology model shows a close fit with the
analytical model for both NLGI 1 and NLGI 2 greases.
Summary and conclusions
4. Investigating the pressure drop in the channel through the
This article presents numerical modeling of lubricating greases in numerical simulations it was found that the values of the
rectangular channels with and without restrictions. The numerical rheological parameters obtained from a plate–plate rheom-
models have been validated with analytical models and compared eter likely are incorrect as reported in Westerberg, et al. (1).
with results from flow visualizations using micro-particle image 5. For the flow in the restricted channels it was found that, in
velocimetry. The main objective of this study is to obtain validated accordance with reported experimental results, the flow
numerical models for flow simulations in more complicated geom- velocity increases as the channel dimensions decrease.
etries where analytical models and flow measurements cannot be Further, the flow depth—that is, the distance the grease
obtained. The particle motion in the grease flow has also been ana- flows between the two double-lip restrictions—was found
lyzed in the channel with two lip restrictions. to be heavily dependent on the rheology of the grease and,
It was found that the simulations match the experiments consequently, considering the shear thinning rheology,
and analysis well, capturing the yield- and shear rate–depen- dependent on the shear rate in the flow.
dent characteristics of lubricating grease flow. The grease flow 6. It was found that particles in the flow are gathered at the
has been modeled as a single-phase Herschel-Bulkley fluid that low-velocity upstream bottom corners of the double-lip
was found to match well with the bulk motion of grease as mea- restrictions.
sured using mPIV.
The findings of the present research work can be summa-
rized as follows: Acknowledgements
1. The CFD model has been validated with a one-dimen- This project is part of Smart Machines and Material (SM2), one of LTU’s
sional analytical model and was also found to be consis- strong research- and innovation areas.
tent with data obtained from measured velocity profiles.
2. It was found that both in restricted and unrestricted
flows the NLGI 00 grease shows a Newtonian-like para- References
bolic flow profile, whereas the NLGI 1 and NLGI 2
(1) Westerberg, L. G., Lundstr€
om, T. S., H€
oglund, E., and Lugt, P. M.
greases show a characteristic non-Newtonian plug flow. (2010), “Investigation of Grease Flow in a Rectangular Channel
The plug length was found to increase with increasing Including Wall Slip Effects Using m;PIV,” Tribology Transactions,
grease thickness (NLGI grade). It was found that a one- 53, pp 600–609.
TRIBOLOGY TRANSACTIONS 13

(2) Barnes, H. A. (2000), “Shear-Thinning Liquids,” A Handbook of Ele- Microparticle Image Velocimetry,” Tribology Transactions, 54,
mentary Rheology, Barnes, H. A. (Ed.), pp 53–61, The University of pp 784–792.
Wales, Institute of Non-Newtonian Fluid, Aberystwyth, UK. (10) Baart, P., Green, T. M., Lundstr€ om, T. S., H€oglund, E., and Lugt, P.
(3) Czarny, R. and Moes, H. (1981), “Some Aspects of Lubricating M. (2011), “The Influence of Speed, Grease Type, and Temperature
Grease Flow,” Proceedings of the 3rd International Congress on Tri- on Radial Contaminant Particle Migration in a Double Restriction
bology, 8(3), pp 68–85. Seal,” Tribology Transactions, 54, pp 867–877.
(4) Czarny, R. (2004), “The Influence of Surface Material and Topography (11) Britton, M. M. and Callaghan, P. T. (1997), “Nuclear Magnetic Reso-
on the Wall Effect of Grease,” Lubrication Science, 14(2), pp 255–273. nance Visualization of Anomalous Flow in Cone-and-Plate Rheome-
(5) Bramhall, A. D. and Hutton, J. F. (1960), “Wall Effect in the Flow of try,” Journal of Rheology, 41, pp 1365–1386.
Lubricating Grease in Plunger Viscosimeters,” British Journal of (12) Magnin, A. and Piau, J. M. (1989), “Application of Freeze–Fracture
Applied Physics, 11, pp 363–369. Technique for Analysing the structure of Lubricating Greases,” Jour-
(6) Vinogradov, G. W., Froishteter, G. B., and Trillski, K. K. (1978), “The nal of Materials Research, 4(4), pp 990–995.
Generalised Theory of Flow of Plastic Disperse Systems with (13) Lugt, P. M. (2013), Grease Lubrication in Rolling Bearings, Wiley &
Account of the Wall Effect,” Rheologica Acta, 17, pp 156–165. Sons, Chichester, UK.
(7) Li, J. X., H€oglund, E., Westerberg, L. G., Green, T. M., Lundstr€
om, T. (14) Kosasih, P. B., Tieu, A. K., and Li, E. B. (2001), “PIV Study of Lubri-
S., Lugt, P. M., and Baart, P. (2012), “mPIV Measurement of Grease cant Flow in Journal Bearing Supply Pocket,” Proceedings of the 14th
Velocity Profiles in Channels with Two Different Types of Flow Australasian Fluid Mechanics Conference, Adelaide University, Ade-
Restrictions,” Tribology International, 54, pp 94–99. laide, Australia, December 10–14.
(8) Scarlett, N. A. (1967), “Use of Grease in Rolling Bearings,” Proceed- (15) COMSOL MultiphysicsÒ v. 5.2. Available at: http://www.comsol.com
ings of the Institution of Mechanical Engineers A, 182, pp 167–171. COMSOL AB, Stockholm, Sweden (accessed January 15, 2016).
(9) Green, T. M., Baart, P., Westerberg, L.-G., Lundstr€ om, T. S., (16) Mitsoulw, E. and Abdali, S. S. (1993), “Flow Simulation of Herschel-
H€oglund, E., Lugt, P. M., and Li, J. (2011), “A New Method to Bulkley Fluids through Extrusion Dies,” The Canadian Journal of
Visualize Grease Flow in a Double Restriction Seal Using Chemical Engineering, 71, pp 147–160.

Você também pode gostar