You are on page 1of 11

David Talbot1

Department of Mechanical
and Aerospace Engineering,
A Helical Gear Pair Pocketing
The Ohio State University,
Columbus, OH 43210
e-mail: talbot.11@osu.edu
Power Loss Model
A new fluid dynamics model is proposed to predict the power losses due to pocketing of
Ahmet Kahraman air, oil, or an air-oil mixture in the helical gear meshes. The proposed computational
Department of Mechanical
procedure treats a helical gear pair as a combination of a number of narrow face width
and Aerospace Engineering,
spur gear segments staggered according to the helix angle and forms a discrete fluid dy-
The Ohio State University,
namics model of the medium being pocketed in the gear mesh. Continuity and conserva-
Columbus, OH 43210
tion of momentum equations are applied to each coupled control volume filled with a
compressible fluid mixture to predict fluid pressure and velocity distributions from which
Satya Seetharaman the instantaneous pocketing power loss is calculated. The proposed model is exercised in
Department of Mechanical order to investigate the fluid pressure and velocity distributions in time along with the
and Aerospace Engineering, pocketing power loss as a function of the speed, helix angle, and oil-to-air ratio.
The Ohio State University, [DOI: 10.1115/1.4026502]
Columbus, OH 43210

1 Introduction of fluid around the gear set, the main components of the spin
power losses of a gear set are: (i) the drag losses of the of the
The efficiency of geared transmission systems has become
rotating components, (ii) losses due to pocketing of the fluid at the
increasingly important due to the demands imposed on automotive
gear mesh interfaces, and (iii) viscous losses of rolling element
and aerospace products in terms of fuel economy. As a result of
bearings, in addition to other secondary sources such as synchron-
the market push and environmental and energy related regulations
izers and oil seals [10].
on vehicles, current and future drive trains and transmissions must
There are various published experimental studies on the power
be designed to meet stringent efficiency requirements. Thus,
loss of a gear or a disk rotating fully, or partially, submerged in
power losses associated with gears have become a major research
lubricant [11–13]. Empirical relations for dimensionless churning
topic in recent years.
torque were provided from experiments for disks [14] and for
As is evident from various published experimental studies
gears in mesh [15,16]. Ariura et al. [17] presented experimental
[1–5], power losses in any geared transmission can be broken into
load independent power loss measurements on jet lubricated spur
two main categories: load-dependent (mechanical) power losses
gears. The experiments by Petry-Johnson et al. [1] provided data
and load-independent (spin) power losses. Load-dependent power
on the influence of the face width and module of gears on spin
losses are attributed to loaded lubricated contacts of the mating
losses under jet lubricated (windage) conditions.
surfaces at the gear meshes and rolling element bearings. They
A theoretical approach to load independent power loss due to
originate from the relative sliding and rolling action of the elasto-
oil fling-off was provided by Akin et al. [18,19]. Models on the
hydrodynamic lubrication (EHL) film at the contact interface that
windage losses of gears rotating only in air have been limited to
can be loosely defined as friction. The state-of-the-art in the mod-
empirical studies [20,21] or computational fluid dynamics models
eling of gear mesh mechanical power losses is represented by a
[22,23]. As stated by the review paper of Eastwick and Johnson
number of recent studies on spur gears (e.g. Refs. [6,7]) where the
[24], most of these studies excluded losses associated with fluid
EHL theory was employed to predict friction at gear contacts.
flow at the gear mesh interface. Pechersky and Wittbrodt [25] pro-
These studies either relied on a real-time transient analysis of the
posed a theoretical analysis to compute the pressure and velocities
EHL conditions of the gear contacts as they move along the tooth
of oil trapped in the meshing zone of spur gears without comput-
surface or used friction models based on the regression analysis of
ing the resultant power loss. Houjoh et al. experimentally investi-
the EHL results covering wide ranges of key contact parameters,
gated the fluid pressure and velocity distributions in spur [26] and
rather than employing any empirical friction formula. Some of
helical [27] gears from the standpoint of improving lubrication
these models included only the sliding power losses under full-
techniques. Diab et al. [21,28] introduced a modeling effort simi-
film lubrication conditions [6], while others captured both sliding
lar to the one proposed here using conservation of mass and con-
and rolling losses that take place at the gear contacts [7] via mixed
servation of energy in order to model the trapping of air only in
(or boundary) EHL formulations [8] to capture actual asperity
the meshing zone, while the use of the conservation of energy to
(metal-to-metal) contacts common to gear interfaces. This EHL-
predict air pressure and velocity for determining power loss is de-
based method was also applied to helical gears by simply treating
batable. Most recently, Seetharaman and Kahraman [10] proposed
them as a collection of narrow face width spur gear slices stag-
a fluid-mechanics based formulation to compute drag power
gered according to the helix angle [9].
losses along the sides and periphery of rotating spur gears and
The mechanisms for spin power losses of gearboxes are entirely
pocketing losses at the gear mesh interface under dip lubrication
different from those of the mechanical losses. Since contact fric-
conditions. This incompressible flow formulation for oil churning
tion is negligible under no-load conditions, interactions of the
losses was later complemented by a compressible flow formula-
gear set components with the surrounding medium (oil, air, or any
tion to predict the same for windage losses under jet lubrication
mixture of oil and air) are responsible for such losses. They have
conditions [29]. The pocketing losses were predicted to represent
often been referred to as churning (in the case of oil) and windage
a large portion of the total windage losses. The same investigators
(in the case of air or an air/oil mix) losses. Regardless of the type
[30] provided comparisons between their predictions and
1
experiments.
Corresponding author.
Contributed by the Tribology Division of ASME for publication in the JOURNAL
Gear pocketing models proposed in Refs. [10,28–30] are either
OF TRIBOLOGY. Manuscript received August 3, 2013; final manuscript received only for spur gears where the pockets formed between the teeth of
January 8, 2014; published online February 24, 2014. Assoc. Editor: Xiaolan Ai. mating gears at any given instant remain constant along the face

Journal of Tribology Copyright V


C 2014 by ASME APRIL 2014, Vol. 136 / 021105-1

Downloaded From: https://tribology.asmedigitalcollection.asme.org on 06/29/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


width direction or for windage only. Since the contact ratio of a
spur gear is lower and the helix angle is zero, a small number of
such pockets must be defined and analyzed for their power losses.
There is no pocketing power loss model available in the literature
for external and internal helical gear pairs operating under dip
(churning) or jet (windage) lubrication conditions.
This paper proposes a new fluid dynamics model to predict the
power loss caused by the pocketing of fluid at a helical gear mesh
interface. The geometries of helical gears are numerically ana-
lyzed in order to provide inputs for the fluid dynamics model in
the form of volume and area values as a function of time. The pro-
posed formulation relies on the prediction of the fluid pressure
and velocity distributions within the gear mesh interface in order
to predict pocketing power loss. While the fluid dynamics formu-
lation proposed here will be presented in a form that is specific to
helical gears (and to spur gears since they are a special case of
helical gears), the fluid dynamics formulations will be generic
such that it may also be used to predict pocketing losses of other
types of gearing such as spiral bevel and hypoid gears. The spe-
cific objectives of this paper can be stated as follows:
(i) Devise a computational methodology to calculate key
time-varying geometrical parameters for external and in-
ternal helical gear meshes, including pocket volumes, exit
areas, and centroids of each pocket.
(ii) Develop a general fluid dynamics model in order to pre-
dict pocketing power losses due to the squeezing of lubri-
cant, air, or a mixture of each, from the gear mesh.
(iii) Exercise the model in order to investigate the effect of the
helix angle and oil-to-air ratio on the fluid pressure and
velocity, along with the pocketing power loss.
2 Model Formulation
2.1 Computation of Helical Gear Mesh Volumes and
Areas. The mesh of a helical gear pair contains a number of pock-
ets that are formed, compressed, and expanded in volume as the
gears rotate in mesh. Figure 1 shows an external spur gear mesh at
different sequential incremental mesh positions to indicate that
there are multiple pockets formed: (i) between the teeth of gear 1
and the root fillets of gear 2, and (ii) between the teeth of gear 2
and the root fillets of gear 1. It is noted that there is a periodicity
to these pockets, i.e., a pocket at a given time instant will have the Fig. 1 Transverse pocket geometry at different mesh positions
same shape as the proceeding pocket that is one base pitch ahead
when the gears are rotated by one mesh cycle.
The cross-sectional shapes shown in Fig. 1 along the transverse
(iv) the IJ centroids of these volumes
plane of the gears remain the same along the face width direction
(v) the ðI þ 1ÞJ circumferential exit areas (including flow
only when the gears are the spur type [10]. In the case of spur
ðnÞ through the gear mesh backlash and the gap between
gears, the volume of a pocket is the product of the side area Aeði;jÞ , active tooth surfaces before and after the gears are in
as shown in Fig. 1, and the active face width of the gears. In the contact)
case of helical gears, however, the cross-sectional shape of the (vi) the ðI þ 1ÞJ centroids of these circumferential exit areas
pocket at a given mesh position changes along the face width
Figure 3(a) illustrates the numerical procedure devised to calcu-
direction according to the rotation transformation defined by the
late the end area of a face width slice j. The minimum distance
helix angle. Figure 2 illustrates a pocket of an external helical
between the two gear surfaces along a slice on both the contacting
gear pair at an arbitrary instantaneous position. The three-
side and the backlash side are found using a search algorithm in
dimensional shape of the pocket formed between a gear tooth and
order to determine the points on both gears that define the pocket.
the space between two teeth of the mating gear is evident in
Grid points are added to either surface in order to have K nodes
Fig. 2(b), where the shape of each transverse slice of the pocket
on both surfaces. This method is based on numerically integrating
along the face width direction varies along with the tooth contact
the area bounded by two mating gear tooth surfaces approximately
points on the instantaneous contact line. Therefore, a form of dis-
by using K  1 quadrilaterals. Figure 3(a) shows a very coarse
cretization is required in the face width direction using the trans-
discretization using four quadrilaterals (K ¼ 5) to demonstrate the
verse involute geometry. For J discrete control volumes (slices) in
procedure. Here the j-th face width slice of the i-th pocket formed
the face width direction and a total of I pockets in the circumfer-
by the mating gears is shown in a rotational increment n. This dis-
ential direction (see Fig. 1), the quantities of interest are:
cretization and the width of the control volume in the face width
(i) the IðJ þ 1Þ end exit areas (to the next control volume or direction yield all of the required pocket information. Among
ðnÞ
to the outside if the slice is the first or the last slice along them, the end area Aeði;jÞ of the face width slice j of a pocket i is
ðnÞ
the face width of the gears) whose normal is in the face given by summing the areas Aði;j;kÞ of each quadrilateral. With the
width direction area of each quadrilateral given as one-half of the magnitude of
(ii) the IðJ þ 1Þ centroids of these end areas the cross product of its diagonals, as shown in Fig. 3(b), the end
(iii) the IJ volumes of discrete control volumes area of the slice is found as

021105-2 / Vol. 136, APRIL 2014 Transactions of the ASME

Downloaded From: https://tribology.asmedigitalcollection.asme.org on 06/29/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 2 (a) Discretization of the helical gear interface into nar-
row gear slices, and (b) a discretized instantaneous helical gear
pocket across the face width

 
K  
 ðxði;j;1;kÞ  xði;j;2;k1Þ Þ 
ðnÞ 1X  
Aeði;jÞ ¼ (1)
2 k¼2 
 ðx x Þ

ði;j;2;kÞ ði;j;1;k1Þ

where the subscript e denotes an end area, the subscript k refers to


the index counting the number of discrete points along the surfa-
ces of pinion (gear index 1) and gear (gear index 2), the subscript
i refers to the meshing pocket in the circumferential direction (see Fig. 3 (a) A coarse discretization of a transverse slice of the
Fig. 1), the subscript j refers to the face width slice being ana- mesh pocket, and (b) a single quadrilateral for area analysis
lyzed, and the superscript n denotes the time step. The vectors x
define the vertices of the kth quadrilateral, as defined in Fig. 3.
ðnÞ ðnÞ With the centroids of the quadrilaterals forming the known end
With this process, the end exit areas Aeði;jÞ and Aeði;jþ1Þ of the jth ðnÞ
area, the overall centroid coordinates xecði;jÞ of the exit area are
discrete control volume of the i-th pocket are determined at the
time increment n. calculated as a weighted sum of the quadrilateral centroids as
ðnÞ
According to Fig. 3(a), the coordinates of the centroids xcði;j;kÞ
of each of the K  1 quadrilaterals are found using any two trian- X
K
ðnÞ ðnÞ
Aði;j;kÞ xecði;j;kÞ
gles that form the quadrilateral as ðnÞ k¼2
xecði;jÞ ¼ ðnÞ
(3)
ðnÞ Ae1 xec1 þ Ae2 xec2 Aeði;jÞ
xcði;j;kÞ ¼ ðnÞ
(2a)
Aði;j;kÞ
The control volume of the pocket defined between the two end
ðnÞ ðnÞ
where areas Aeði;jÞ and Aeði;jþ1Þ is approximately defined as

1 h ðnÞ i ðnÞ 1 h ðnÞ ðnÞ


i
xec1 ¼
ðnÞ ðnÞ
xði;j;1;k1Þ þ xði;j;2;kÞ þ xði;j;2;k1Þ (2b) Vði;jÞ ¼ Db Aeði;jÞ þ Aeði;jþ1Þ (4)
3 2

1 h ðnÞ ðnÞ ðnÞ


i where Db is the face width of the jth control volume of the ith
xec2 ¼ xði;j;1;k1Þ þ xði;j;1;kÞ þ xði;j;2;kÞ (2c) ðnÞ
3 pocket being analyzed. Likewise, the centroid xcði;jÞ of this control
  volume is the average of the centroids of its end exit areas
 ðnÞ ðnÞ 
 ðxði;j;2;kÞ  xði;j;1;k1Þ Þ 
1
Ae1 ¼  

 (2d) ðnÞ 1 h ðnÞ ðnÞ
i
2  ðnÞ ðnÞ  xcði;jÞ ¼ xecði;jÞ þ xecði;jþ1Þ (5)
 ðxði;j;2;k1Þ  xði;j;1;k1Þ Þ  2
ðnÞ ðnÞ
  The circumferential exit areas Arði;jÞ and Arðiþ1;jÞ are calculated
 ðnÞ ðnÞ  ðnÞ
 ðxði;j;2;kÞ  xði;j;1;k1Þ Þ  from the backlash distance drði;jÞ ¼ jjxði;j;2;k¼1Þ  xði;j;1;k¼1Þ jj and
1 
Ae2 ¼    (2e) ðnÞ
the contact gap distance drðiþ1;jÞ ¼ jjxði;j;2;k¼K Þ  xði;j;1;k¼K Þ jj of the
2  ðnÞ ðnÞ 
 ðxði;j;1;kÞ  xði;j;1;k1Þ Þ 
transverse slice, as shown in Fig. 3(a). They are given as

Journal of Tribology APRIL 2014, Vol. 136 / 021105-3

Downloaded From: https://tribology.asmedigitalcollection.asme.org on 06/29/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


ðnÞ Db h ðnÞ ðnÞ
i
Arði;jÞ ¼ drði;jÞ þ drði;jþ1Þ (6)
2
ðnÞ
The centroid coordinates xrcði;jÞ of the circumferential exit area are
calculated in the same manner as the end exit area individual
quadrilateral centroids in Eq. (2) with k ¼ 1 or k ¼ K, depending
whether the calculation is on the contact or backlash side as
ðnÞ Ar1 xrc1 þ Ar2 xrc2
xrcði;jÞ ¼ ðnÞ
(7a)
Arði;jÞ
where

1 h ðnÞ ðnÞ ðnÞ


i
xrc1 ¼ xði;j;1;kÞ þ xði;jþ1;1;kÞ þ xði;jþ1;2;kÞ (7b)
3

1 h ðnÞ ðnÞ ðnÞ


i
xrc2 ¼ xði;j;1;kÞ þ xði;j;2;kÞ þ xði;jþ1;2;kÞ (7c)
3
 
 ðnÞ ðnÞ 
ðxði;jþ1;1;kÞ  xði;j;1;kÞ Þ 
1 
Ar1 ¼    (7d)
2  ðnÞ ðnÞ 
 ðxði;jþ1;2;kÞ  xði;j;1;kÞ Þ 
 
 ðnÞ ðnÞ 
 ðxði;j;2;kÞ  xði;j;1;kÞ Þ 
1 
Ar2 ¼    (7e)
2  ðnÞ ðnÞ 
 ðxði;jþ1;2;kÞ  xði;j;1;kÞ Þ 

This process is applied to the transverse slices of each tooth, be-


ginning when a tooth first impinges upon a pocket (the cavity
between the two teeth of the mating gear) and ending when the
tooth completely leaves that pocket. Repeating this for a discrete
number of time steps through the mesh cycle, the time variations
of each control volume and its centroid, along with the escape
areas and centroids of the escape areas associated with this control
volume, are determined. A sample calculation is presented in
ðnÞ
Fig. 4 to illustrate the time variation of an end area Aeði;jÞ ,
ðnÞ ðnÞ
backlash-side area Arði;jÞ , and contact-side area Arðiþ1;jÞ of a con-
trol volume of an external gear mesh through the meshing cycle.
The time scale used in this figure is mesh cycles # that represent
gear rotation normalized with respect to the base pitch (i.e.,
# ¼ h=hbp where h is the gear rotation angle and hbp ¼ 2p=Z;
here, Z is the number of teeth of the same gear). The areas in
Fig. 4 are plotted as a function of #. It is noted in Fig. 4(a) that
ðnÞ
Aeði;jÞ reduces in a nearly quadratic manner up to a certain rota-
tional position where it reaches its minimum before quadratically
expanding again. This indicates that the control volume associated
with this end area first shrinks and then expands. A similar varia- Fig. 4 An example variation of (a) the end area, (b) the
tion is observed for the backlash-side area in Fig. 4(b), where a backlash-side area, and (c) the contact-side area of the control
sizable portion of the meshing cycle is represented by a small but volume j of a pocket i with the mesh position # for b 5 0
ðnÞ
constant Arði;jÞ defined by the nominal backlash of the gear pair. cross in Fig. 5, indicates that these two pockets are separated by
Likewise, the contact-side escape area also varies in a similar the tooth contact, while an open connection represents a passage
manner, except there is a segment of the cycle with zero area, along the backlash (or gap between the mating teeth) allowing
indicating that this opening has been completely blocked by the transport of the media between the pockets. The columns of pock-
tooth contact. ets in Fig. 5 indicate J transverse control volumes of each face
With the pocket sizes and escape areas established, along with width slice that are connected to each other by the end areas
the variations of these quantities with time (or #), the problem at defined in Eq. (1). All of the IJ control volumes shown in Fig. 5
hand reduces to a multidegree-of-freedom discrete fluid dynamics change their volumes and areas of their openings to ambient or
problem, as shown in Fig. 5. Here, multiple time-varying control other control volumes as the gears roll.
volumes are connected to each other in the circumferential direc-
tion through backlash and contact gap areas and, in the face width
direction, through the end areas. In the vertical direction, a total of 2.2 External and Internal Helical Gear Mesh Pocketing
I pockets exist along the circumference of the gears with or Power Loss Formulation. In Fig. 5, we consider the fluid flow
without connections to the adjacent pockets. Here the number of into and from any given discrete control volume j of pocket i. The
pockets I is not constant and varies with #, also making the continuity equation for this control volume and the conservation
degrees of freedom variable. A blocked connection, denoted by a of momentum (ignoring viscous effects and speeds approaching

021105-4 / Vol. 136, APRIL 2014 Transactions of the ASME

Downloaded From: https://tribology.asmedigitalcollection.asme.org on 06/29/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 5 Multi-degree-of-freedom fluid dynamics model governing the helical gear pocketing
problem

  
the speed of sound) at the exits of the control volume are given, ðnÞ ðnÞ  ðnÞ ðnÞ 
pði;jÞ  pði1;jÞ xrcði;jÞ  xcði1;jÞ 
respectively, as ðnÞ ðnÞ
prði;jÞ ¼ pði1;jÞ þ 
 ðnÞ
   (11)
ðnÞ   ðnÞ ðnÞ 
ð ð xrcði;jÞ  xcði;jÞ  þ xrcði;jÞ  xcði1;jÞ 
@
qdV þ qudA ¼ 0 (8)
@t CV CS With the density at the axial connections qeði;jÞ and circumferential
ðnÞ

@ @p ðnÞ
connections qrði;jÞ defined, the discrete form of Eq. (8) is given as
ðquÞ þ ¼0 (9)
@t @x 0 1 0 1
ðnÞ ðn1Þ ðnÞ ðn1Þ
V  Vði;jÞ
ðnÞ @ ði;jÞ ðnÞ
qði;jÞ  qði;jÞ
where q is the fluid density, V is the pocket volume, A is the exit qði;jÞ AþV @ A
Dt ði;jÞ Dt
area, u is the exit velocity and @p=@x is the pressure gradient
across an exit. In these equations, CV denotes integration over the
X
1 X
1
control volume and CS denotes integration over the exit surfaces þ
ðnÞ
1k qeðiþk;jÞ Aeðiþk;jÞ ueðiþk;jÞ þ
ðnÞ
1k qrði;jþkÞ Arði;jþkÞ urði;jþkÞ ¼ 0
of the control volume. Considering the calculated geometrical k¼0 k¼0
inputs and the index notation presented in Sec. 2.1 for a discrete
(12)
number of IJ control volumes shown in Fig. 5, while also assum-
ing pressure and velocity gradients between the pockets to be lin-
ear, Eqs. (8) and (9) can be applied in discretized form for both where Dt is the time step that is dependent on the rotational speed
space and time. and the number of teeth of the driving gear (gear 1). Here,
1k ¼ 1 for k ¼ 0 and 1k ¼ 1 for k ¼ 1. Discretized conservation
Consider the jth control volume of the i-th circumferential
ðnÞ ðnÞ of momentum (see Eq. (9)) equations in the circumferential and
pocket with pressure pði;jÞ and density qði;jÞ . The circumferential axial directions are given, respectively, by
ðnÞ
fluid velocity urði;jÞ is the fluid velocity between the axial control 0 1 0 1
ðnÞ ðn1Þ ðnÞ ðn1Þ
volumes j of pockets i and i  1. Axial fluid velocity ueði;jÞ is the
ðnÞ
ðnÞ
urði;jÞ  urði;jÞ ðnÞ
qrði;jÞ  qrði;jÞ
qrði;jÞ @ Aþ urði;jÞ @ A
fluid velocity between pocket i of the axial control volumes j and Dt Dt
j  1. Assuming a linear pressure gradient between control vol- 0 1
ðnÞ ðnÞ ðnÞ
umes, the pressure peði;jÞ at the axial connections and the pressure ðnÞ ðnÞ B
urðiþ1;jÞ  urði1;jÞ C
ðnÞ þ 
qrði;jÞ urði;jÞ @ A
prði;jÞ at the circumferential connections are defined as ðnÞ ðnÞ 
xrcðiþ1;jÞ  xrcði1;jÞ 
   0 1
ðnÞ ðnÞ  ðnÞ ðnÞ  ðnÞ ðnÞ
pði;jÞ  pði;j1Þ xecði;jÞ  xcði;j1Þ  p  p
ðnÞ ðnÞ
    B ði;jÞ ði1;jÞ
C
peði;jÞ ¼ pði;j1Þ þ  (10) þ @  
A ¼ 0; (13)
ðnÞ ðnÞ   ðnÞ ðnÞ   ðnÞ ðnÞ   ðnÞ ðnÞ
xecði;jÞ  xcði;jÞ  þ xecði;jÞ  xcði;j1Þ  xrcði;jÞ  xcði;jÞ  þ xrcði;jÞ  xcði1;jÞ 

Journal of Tribology APRIL 2014, Vol. 136 / 021105-5

Downloaded From: https://tribology.asmedigitalcollection.asme.org on 06/29/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


0 1 0 1
ðnÞ ðn1Þ ðnÞ ðn1Þ X
I X
J
ueði;jÞ  ueði;jÞ qeði;jÞ  qeði;jÞ ðnÞ
ðnÞ
qeði;jÞ @ A þ uðnÞ @
eði;jÞ
A PpðnÞ ¼ W_ ði;jÞ (17a)
Dt Dt i¼1 j¼1
0 1
ðnÞ ðnÞ
ðnÞ ðnÞ B
u eði;jþ1Þ  u eði;j1Þ C Finally, the average pocketing power loss associated with the
þ qeði;jÞ ueði;jÞ @
 ðnÞ
A
 meshing of two helical gears is given as
ðnÞ
xecði;jþ1Þ  xecði;j1Þ 
0 1 1X N
ðnÞ ðnÞ
pði;jÞ  pði;j1Þ Pp ¼ PðnÞ (17b)
B C N n¼1 p
þ @ ðnÞ
 
A ¼ 0 (14)
ðnÞ   ðnÞ ðnÞ
xecði;jÞ  xcði;jÞ  þ xecði;jÞ  xcði;j1Þ 
To this point, the modeling has been done by assuming that the
density of the fluid in the mesh is defined. These formulations are
Here, the exits that are considered to be at the ambient pressure significantly simplified for a gear mesh completely filled with an
include the circumferential exits of index i ¼ 1 or I þ 1, axial end incompressible lubricant. As the other extreme, for a mesh com-
exits of index j ¼ 1 or J þ 1, and any exits connecting a control pletely filled with air, the ideal gas law qa ¼ p=ðRTÞ, can be
volume being analyzed to a control volume left out of the analysis employed for the density-pressure relationship where R is the spe-
(control volumes with no tooth impinging on the cavity). For these cific gas constant for air. For the most common and practical case
cases, the pressure gradient term in the momentum equations are of an air-oil mixture, an equivalent density of the air-oil mixture
modified accordingly. is defined by using the volumetric lubricant-to-air ratio n at ambi-
If there are I circumferential pockets and J control volumes in ent pressure. Furthermore, the density of the lubricant in the mix-
the face width direction, this set provides IJ þ IðJ þ 1Þ þJðI þ 1Þ ture is assumed to remain constant (i.e., act as an incompressible
equations in order to solve for the IJ pocket pressures, JðI þ 1Þ fluid) while the air in the mixture is allowed to expand (be com-
circumferential velocities, and IðJ þ 1Þ axial velocities for any pressible), according to the ideal gas law, yielding an approximate
time step n, given the initial conditions pðn1Þ and uðn1Þ . New- equivalent density
ton’s method is used to solve the set of nonlinear equations for
each time step individually until one mesh cycle is completed and  
pa
the final mesh position solution is used as an initial condition for nqo þ ð1  nÞ
RT
the beginning of the mesh cycle. This process is continued until qeq ¼  a  (18)
the solution approaches a steady state solution within an allowable pa T
n þ ð1  nÞ Ta
error threshold. p
Using the preceding solution for the pocket pressures and exit
velocities, the conservation of energy for the control volumes Here, qo is the lubricant (oil) density and pa and Ta are the ambi-
(ignoring gravitational effects) yields ent pressure and temperature. The temperature T of the air in each
ð ð   pocket is also required in order to define qeq , such that IJ relations
@ p ðnÞ
ðqEÞdV þ Au þ E qdA ¼ W_ (15) are needed for the unknowns Tði;jÞ . Assuming isentropic expansion
@t CV CS q
of the air
where E ¼ cv T þ 12u2 is the internal energy of the fluid, T is the 0 1ðc1=cÞ
ðnÞ ðnÞ
temperature, cv is the constant volume specific heat, and W_ is the Tði;jÞ pði;jÞ
work done on the control volume that is equal to the pocketing ðn1Þ
¼@ ðn1Þ
A (19)
Tði;jÞ pði;jÞ
power loss.
Assuming uniform flow at the entrances from ambient condi-
tions (ak ¼ 1) and fully developed flow elsewhere (ak ¼ 2), the where c is the specific heat ratio of air. This adds IJ equations to
same discretization scheme used earlier is applied to Eq. (15) in the solution scheme for a complete solution.
order to find the power loss associated with pocketing of the jth
control volume of the ith circumferential pocket as
3 Results and Discussion
2 3 Example pocketing power loss simulations of external helical
ðnÞ ðn1Þ gears are presented in this section while the proposed methodol-
ðnÞ
ðqVT Þði;jÞ ðqVT Þði;jÞ
W_ ði;jÞ ¼ cv 4 5 ogy also applies to internal helical gear meshes. A family of four
Dt unity-ratio external gear pairs is considered containing the exam-
2 3 ple gear sets. All four gear sets have the same transverse geome-
ðnÞ ðn1Þ
ðqVu2 Þði;jÞ ðqVu2 Þði;jÞ try, but different helix angles of b ¼ 0 deg; 5 deg, 15 deg, and
þ4 5
30 deg. Table 1 lists the geometric parameters of these example
Dt
2   3 gear pairs. All of the analyses presented here are performed with a
X1
ðnÞ ðnÞ ðnÞ
cv qrðiþk;jÞ Trðiþk;jÞ þ prðiþk;jÞ lubricant having a density of 824 kg=m3 and a temperature of
1 ðnÞ ðnÞ 6 7 Ta ¼ 90 C. A total of J ¼ 21 control volumes were used
þ A u 4  2 5
2 k¼0 rðiþk;jÞ rðiþk;jÞ ðnÞ ðnÞ (j 2 ½1; 21) such that each axial gear slice has a width
þak qrðiþk;jÞ urðiþk;jÞ
2   3 Db ¼ b=J
 ¼ 1:19 mm for the given gear face width of b ¼ 25
ðnÞ ðnÞ ðnÞ mm. A dimensionless axial position parameter b is defined as
X1 cv qeði;jþkÞ Teði;jþkÞ þ peði;jþkÞ
1 ðnÞ ðnÞ 6 7 b ¼ z=b, where z 2 ½0; b.
 With this, the side boundaries of the
þ A u 4  2 5
2 k¼0 eði;jþkÞ eði;jþkÞ þa qðnÞ ðnÞ first control volume j ¼ 1 are b ¼ 0:0 and 0.048 and the side
k eði;jþkÞ ueði;jþkÞ
boundaries of the last control volume j ¼ J ¼ 21 are at b ¼ 0:952
(16) and 1.0. Likewise, the centers of the control volumes j ¼ 1; 11,
and 21 are b ¼ 0:024; 0.5, and 0.976, respectively.
Figure 6(a) shows the time variation of the fluid pressure pði;jÞ
Summing up the individual control volume power losses at a of pocket i at b ¼ 0:5 of the spur gear pair (b ¼ 0) rotating at
given position n, the total pocketing loss of the helical gear mesh X ¼ 3000 rpm in a fluid mixture of n ¼ 0:05. Meanwhile,
in this instantaneous position is found as Fig. 6(b) shows the time histories of the end exit velocities ueði; jÞ

021105-6 / Vol. 136, APRIL 2014 Transactions of the ASME

Downloaded From: https://tribology.asmedigitalcollection.asme.org on 06/29/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Table 1 Parameters of example external helical gear sets values. The analysis (pocketing action) of this pocket i is shown
to begin at # ¼ 0:60; where the tip of one gear penetrates the out-
Gear 1 Gear 2 side diameter of pocket i being analyzed and continues for
approximately three base pitches of rotation. Between # ¼ 0:60
Number of teeth 25 25 and 1.40, a rise in fluid pressure is predicted in Fig. 6(a) that is
Transverse module (mm) 2
Helix angle (deg) 0, 5, 15, 30
accompanied by an acceleration of the fluid out of the pocket, as
Transverse pressure angle (deg) 20 shown in Fig. 6(b). Shortly following this, the fluid is rapidly
Outside diameter (mm) 54.00 54.00 decelerated as the pressure quickly drops. At # ¼ 2:05, pocket i
Root diameter(mm) 44.60 44.60 stops compressing and begins expanding such that the end exit
Center distance (mm) 50.00 velocities in Fig. 6(b) are instantaneously zero (ueði; jÞ ¼ 0 at both
Face width (mm) 25.00 25.00 b ¼ 0:0 and 1.0). In Fig. 6(b), the two exit velocities have the
Transverse tooth thickness (mm) 3.0916 3.0916 same magnitude and opposite directions since b ¼ 0. It is noted
Diameter of thickness meas. (mm) 50.00 50.00 here that the pressure profile of Fig. 6(a) qualitatively matches
Transverse backlash (mm) 0.10
well with the measurements of Diab et al. [28], showing the com-
pression and expansion cycle with a larger absolute gauge pres-
sure magnitude in the negative pressure portion of the cycle. This
of the same circumferential pocket i at b ¼ 0:0 and 1.0. Likewise, is required for the rapid deceleration of the fluid. The model also
the time histories of the contact- and backlash-side circumferen- captures a pressure ripple at the end of the cycle (at # ¼ 3:35)
tial exit velocities urði; jÞ and urðiþ1; jÞ of the same control volume caused by the re-opening of the contact side area. Such a ripple is
of pocket i at b ¼ 0:5 are shown in Figs. 6(c) and 6(d). In these also evident in the measurements of Diab et al. [28].
figures, certain rotational positions at # ¼ 0:60; 1.40, 2.05, and As is seen from the time history of the contact-side exit velocity
3.35 are marked by vertical dotted lines. The corresponding posi- in Fig. 6(c), the fluid is allowed to backflow from pocket i while it
tions of pocket i in the meshing are shown in Fig. 7 at these four # is being compressed into the previous pocket i  1 until the gears

Fig. 6 (a) Pressure, (b) end velocity, (c) contact-side velocity, and (d) backlash-side velocity
time histories for the example spur gear pair (b 5 0 deg) at X 5 3000 rpm and n 5 0:05

Fig. 7 Side views of the spur gear pair of Fig. 6 at representative mesh positions. The pocket i
of interest is marked with .

Journal of Tribology APRIL 2014, Vol. 136 / 021105-7

Downloaded From: https://tribology.asmedigitalcollection.asme.org on 06/29/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 8 Pressure and exit velocity time histories of the example spur gear pair at X 5 4000 rpm
for (a) n 5 0:01 and (b) n 5 0:05

Fig. 9 End exit velocity time histories of gear pairs with helix angles of (a) b 5 0 deg, (b)
b 5 5 deg, (c) b 5 15 deg, and (d) b 5 30 deg at X 5 3000 rpm and n 5 0:05

come into contact to block the passage (Arði;jÞ ¼ 0 in Fig. 4(c)). After # ¼ 1:40, the connected pocket ahead begins expanding
The fluid flow through the contact-side exit area is abruptly inter- while the pocket i being analyzed is compressed. Thus, a rapid
rupted at # ¼ 1:40 by the tooth pair coming into contact. flow forward out of pocket i is observed within # 2 ½1:40; 1:90 in
urði; jÞ ¼ 0 at b ¼ 0:5 within # 2 ½1:40; 3:35 represents the period Fig. 6(d). Beyond # ¼ 1:90, pocket i expands more rapidly to
when the contact is maintained. Once the contact is lost at cause the fluid to backflow again into the pocket.
# ¼ 3:35, a reverse flow from pocket i  1 is predicted as pocket i The influence of the oil-to-air ratio on the fluid pressures and
is rapidly expanding. Meanwhile, as shown in Fig. 6(d), the fluid end exit velocities are shown in Fig. 8 using the same spur gear
initially flows backward into the pocket i through the backlash pair as in Fig. 6. In Fig. 8(a), the pði;jÞ and ueði;jÞ time histories are
area from the pocket ahead of it that is being rapidly compressed. shown for pocket i at different b for n ¼ 0:01 when the gear

021105-8 / Vol. 136, APRIL 2014 Transactions of the ASME

Downloaded From: https://tribology.asmedigitalcollection.asme.org on 06/29/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 10 Pressure time histories of gear pairs with helix angles of (a) b 5 0 deg, (b) b 5 5 deg, (c)
b 5 15 deg, and (d) b 5 30 deg at X 5 3000 rpm and n 5 0:05

Fig. 11 End exit velocity distributions of gear pairs with helix angles of (a) b 5 0 deg, (b)
b 5 5 deg, (c) b 5 15 deg, and (d) b 5 30 deg at X 5 3000 rpm and n 5 0:05

operates at X ¼ 4000 rpm. Figure 8(b) shows the same for for the significant differences in the pði;jÞ time histories. The flow
n ¼ 0:05. Comparing Figs. 8(a) and 8(b), only slight differences of fluid at n ¼ 0:05 carries much more momentum than that of the
in the ueði;jÞ profiles are observed between the cases of n ¼ 0:01 flow at n ¼ 0:01, requiring a much larger pressure gradient in
and n ¼ 0:05, while the same cannot be said for pði;jÞ . Since the order to accelerate and decelerate the fluid. This difference in
geometry and rotational speed of the system is the same between behavior points to the need to include the lubricant density in the
the two analyses, the volumetric flow rates between the two differ analysis.
only slightly due to compressibility. This is evident in the fluid ve- Since the main focus of this study is on the pocketing losses of
locity time histories. On the contrary, the difference in the density helical gears, the pocket pressure and velocity distributions of
of the medium between n ¼ 0:01 and n ¼ 0:05 is the main reason gears having helix angles b ¼ 0 deg; 5 deg; 15 deg, and 30 deg will

Journal of Tribology APRIL 2014, Vol. 136 / 021105-9

Downloaded From: https://tribology.asmedigitalcollection.asme.org on 06/29/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


influence of the helix angle and the face width of the gear pair on
pocketing losses. With this modest face width, the b ¼ 30 deg
gear pair shows a sizable reduction in pocketing power loss, while
the b ¼ 15 deg gear pair does not.
Figure 12(b) shows the influence of the oil-to-air ratio n on Pp
of the helical gear pair having b ¼ 30 deg. The Pp values corre-
sponding to n ¼ 0:01, 0.05, 0.15, and 0.25 are compared within a
speed range of 0 to 5000 rpm. Here, higher n values are shown to
result in a higher Pp . As shown in Fig. 12(a), an exponential rela-
tionship is evident between Pp and X in the form of Pp ¼ c Xd ,
where d ¼ 3:01; 3.18, 3.20, and 3.26 for n ¼ 0:01; 0.05, 0.15, and
0.25, respectively. This rather large variation in Pp with n is a
direct result of the differing pressure distributions illustrated in
Fig. 8 and the effective density of the fluid medium filling the
meshing zone. This is also evident from Eq. (16) in terms of the
pressure and effective density.

4 Conclusions
Pocketing power losses can contribute a significant amount to
the overall power loss of a power transmission system, especially
under lightly loaded and high speed applications. Wide face
widths and small helix angles show larger pocketing power loss
effects. In order to quantify this effect, a new fluid dynamics
model was proposed in this study in order to predict the power
losses of helical gear meshes due to the pocketing of air, oil, or an
air-oil mixture in helical gear meshes. This model treated a helical
gear pair as a combination of a number of narrow face width spur
gear slices staggered according to the helix angle. With each cir-
cumferential pocket discretized in this manner, the time variation
of the volumes and escape areas associated with each control vol-
ume were defined to form a discrete fluid dynamics model of a
Fig. 12 Variation of pocketing power loss with the (a) helix
angle at n 5 0:05, and (b) oil-to-air ratio at b 5 30 deg medium being pocketed in the gear mesh. Continuity and conser-
vation of momentum equations were applied to each coupled con-
trol volume filled will a compressible fluid mixture to predict the
be compared next in Figs. 9 and 10 at X ¼ 3000 rpm and fluid pressure and velocity distributions from which the instanta-
n ¼ 0:05. As stated earlier, the transverse geometries of these neous pocketing power loss was calculated. This proposed model
gears (as listed in Table 1) are identical such that b would be the was used to investigate the mechanisms for the helix angle and
sole reason for the differences in the results. The dominant influ- the air-to-oil ratio to influence the pocket pressure and velocity
ence of b on the axial fluid velocities is demonstrated in Fig. 9. As distributions. The impact of the same parameters on the helical
b increases, fluid is driven more and more out of one side of the gear pocketing power losses was also quantified within the ranges
gear pocket. The spur gear pair shows symmetric exiting veloc- of the operating speed.
ities with respect to the midplane of the gears at b ¼ 0:5. With an The complexity of the problem, including rapidly changing
increase in b, this becomes more skewed to drive all of the fluid geometries, has limited most recent computational fluid dynamics
out one side of the gear pair. Larger helix angles allow for pro- studies of gears to single gear analyses [22,23], completely avoid-
gressively larger end and circumferential exit areas across the face ing pocketing losses. The computational model presented here
width, decreasing the magnitudes of the exit velocities in the pro- provides a relatively fast method of predicting power loss due to
cess. Because the maximum exit velocities are decreased, the nec- pocketing of an air/oil mixture from a helical gear mesh. The anal-
essary change in the momentum of the fluid is decreased. yses presented here took approximately 1–3 min using one
Therefore, significantly lower pocket pressure deviations are 3.4 GHz processor, depending upon the considered operation and
experienced in helical gear pairs with larger b, as shown in fluid conditions. Thus, the proposed model can also be used in
Fig. 10. These pressure distribution results also qualitatively agree design evaluations. Finally, the computational methodology pre-
well with the measurements of Diab et al. [28]. sented in this work is also suitable for other gear types. The
Figures 9 and 10 indicate that helical gears cannot be assumed authors have recently applied this formulation to spiral bevel and
to behave like spur gears in terms of their pocketing power losses. hypoid gears, which will be presented in a separate paper.
This variance in the face width direction can be shown more
clearly by plotting the data of Fig. 9 as a function of b at various
discrete rotational positions #, as shown in Fig. 11. Here, the
Acknowledgment
effect of b in driving the flow in one direction is evident. The The authors would like to thank General Motors Powertrain
symmetric flow (squeezing in both directions) seen in Fig. 11(a) is and, specifically, Dr. Avinash Singh for motivating and supporting
progressively changed to mainly flow in one direction denoted by this research activity.
the positive fluid velocity across the face width in
Figs. 11(b)–11(d). References
Figure 12(a) shows the variation of the resultant pocketing [1] Petry-Johnson, T., Kahraman, A., Anderson, N. E., and Chase, D. R., 2008,
power loss Pp of gear pairs having b ¼ 0 deg; 15 deg, and 30 deg “An Experimental Investigation of Spur Gear Efficiency,” ASME J. Mech.
with the speed at n ¼ 0:05. Here, the spur gear (b ¼ 0 deg) is pre- Des., 130, p. 062601.
dicted to have Pp values that are almost three times those for the [2] Yada, T., 1972, “The Measurement of Gear Mesh Friction Losses,” ASME
Paper No. 72-PTG-35, pp. 8–12.
helical gear pair with b ¼ 30 deg at X ¼ 10; 000 rpm while the [3] Naruse, C., Haizuka, S., Nemoto, R., and Kurokawa, K., 1986, “Studies on Fric-
difference between gear pairs having b ¼ 0 deg and 15 deg is tional Loss, Temperature Rise and Limiting Load for Scoring of Spur Gear,”
about 10%. This demonstrates the complex and combined Bull. JSME, 29(248), pp. 600–608.

021105-10 / Vol. 136, APRIL 2014 Transactions of the ASME

Downloaded From: https://tribology.asmedigitalcollection.asme.org on 06/29/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use


[4] Naruse, C., Haizuka, S., Nemoto, R., and Takahashi, H., 1991, “Influences of [18] Akin, L. S. and Mross, J. J., 1975, “Theory for the Effect of Windage on the
Tooth Profile on Frictional Loss and Scoring Strength in the Case of Spur Lubricant Flow in the Tooth Spaces of Spur Gears,” ASME J. Eng. Ind., 97, pp.
Gears,” Proceedings of MPT’’91, JSME International Conference on Motion 1266–1273.
and Power Transmissions, Hiroshima, Japan. [19] Akin, L. S., Townsend, J. P., and Mross, J. J., 1975, “Study of Lubricant Jet
[5] Mizutani, H. and Isikawa, Y., 1996, “Power Loss of Long Addendum Spur Gears,” Flow Phenomenon in Spur Gears,” ASME J. Lubr. Technol., 97, pp. 288–295.
International Conference on Gears, Dresden, Germany, April 22–24,1996, VDI [20] Dawson, P. H., 1984, “Windage Loss in Larger High-Speed Gears,” Proc. Inst.
BERICHTE, 1230, pp. 83–95. Mech. Eng., Part A, 198(1), p. 51–59.
[6] Xu, H., Kahraman, A., Anderson, N. E., and Maddock, D., 2007, “Prediction of [21] Diab, Y., Ville, F., and Velex, P., 2006, “Investigations on Power Losses in High
Mechanical Efficiency of Parallel-Axis Gear Pairs,” ASME J. Mech. Des., Speed Gears,” Proc. Inst. Mech. Eng., Part J: J. Eng. Tribol., 220, pp. 191–298.
129(1), pp. 58–68. [22] Wild, P. M., Dijlali, N., and Vickers, G. W., 1996, “Experimental and Compu-
[7] Li, S. and Kahraman, A., 2010, “Prediction of Spur Gear Mechanical Power tational Assessment of Windage Losses in Rotating Machinery,” ASME Trans.
Losses Using a Transient Elastohydrodynamic Lubrication Model,” Tribol. J. Fluids Eng., 118, pp. 116–122.
Trans., 53, pp. 554–563. [23] Al-Shibl, K., Simmons, K., and Eastwick, C. N., 2007, “Modeling Gear Wind-
[8] Li, S. and Kahraman, A., 2010, “A Transient Mixed Elastohydrodynamic Lubri- age Power Loss From Enclosed Spur Gears,” Proc. Inst. Mech. Eng., Part A,
cation Model for Spur Gear Pairs,” ASME J. Tribol., 132(1), p. 011501. 221(3), pp. 331–341.
[9] Li, S., Vaidyanathan, A., Harianto, J., and Kahraman, A., 2009, “Influence of [24] Eastwick, C. N. and Johnson, G., 2008, “Gear Windage: A Review,” ASME J.
Design Parameters on Mechanical Power Losses of Helical Gear Pairs,” JSME Mech. Des., 130(3), p. 034001.
J. Adv. Mech. Des. Syst. Manuf., 3(2), pp. 146–158. [25] Pechersky, M. J. and Wittbrodt, M. J., 1989, “An Analysis of Fluid
[10] Seetharaman, S. and Kahraman, A., 2009, “Load-Independent Power Losses of Flow Between Meshing Spur Gear Teeth,” Proceedings of the ASME Fifth
a Spur Gear Pair: Model Formulation,” ASME J. Tribol., 131(2), p. 022201 International Power Transmission and Gearing Conference, Chicago, IL, pp.
[11] Changenet, C. and Velex, P., 2007, “A Model for the Prediction of Churning 335–342.
Losses in Geared Transmissions—Preliminary Results,” ASME J. Mech. Des., [26] Houjoh, H., Ohshima, S., Miyata, S., Takimoto, T., and Maenami, K., 2000,
129, pp. 128–133. “Dynamic Behavior of Atmosphere in a Tooth Space of a Spur Gear During
[12] Daily, J. and Nece, R., 1960, “Chamber Dimension Effects of Induced Flow Mesh Process From the Viewpoint of Efficient Lubrication,” Proceedings of the
and Frictional Resistance of Enclosed Rotating Disks,” ASME J. Basic Eng., ASME, Design Engineering Technical Conference, Baltimore, MD, Paper No.
82, pp. 217–232. PTG-14372.
[13] Mann, R. and Marston, C., 1961, “Friction Drag on Bladed Disks in Housings [27] Houjoh, H., Ohshima, S., Matsumura, S., Yumia, Y., and Itoh, K., 2003,
as a Function of Reynolds Number, Axial and Radial Clearance, and Blade As- “Pressure Measurement of Ambient Air in the Root Space of Helical Gears for
pect Ratio and Solidity,” ASME J. Basic Eng., 83, pp. 719–723. the Purpose of Understanding Fluid Flow to Improve Lubrication Efficiency,”
[14] Boness, R. J., 1989, “Churning Losses of Discs and Gear Running Partially Proceedings of the ASME, Design Engineering Technical Conference, Chicago,
Submerged in Oil,” Proceedings of the ASME International Power Transmis- IL, Paper No. PTG-48117.
sion and Gearing Conference, Chicago, IL, pp. 355–359. [28] Diab, Y., Ville, F., Houjoh, H., Sainsot, P., and Velex, P., 2005, “Experimental
[15] Luke, P. and Olver, A. V., 1999, “A Study of Churning Losses in Dip- and Numerical Investigations on the Air-Pumping Phenomenon in High-Speed
Lubricated Spur Gears,” Proc. Inst. Mech. Eng., Part G, 213, pp. 337–346. Spur and Helical Gears,” Proc. Inst. Mech. Eng., Part C, 219, pp. 785–800.
[16] Terekhov, A. S., 1991, “Basic Problems of Heat Calculation of Gear Reducers,” [29] Seetharaman, S. and Kahraman, A., 2010, “A Windage Power Loss Model for
Proceedings of the JSME International Conference on Motion and Power Trans- Spur Gear Pairs,” Tribol. Trans., 53(4), pp. 473–484.
missions, pp. 490–495. [30] Seetharaman, S., Kahraman, A., Moorhead, M, and Petry-Johnson, T., 2009,
[17] Ariura, Y., Ueno, T., and Sunaga, T., 1973, “The Lubricant Churning Loss in “Oil Churning Power Losses of a Gear Pair: Experiments and Model Vali-
Spur Gear Systems,” Bull. JSME, 16, pp. 881–890. dation,” ASME J. Tribol., 131, p. 022202.

Journal of Tribology APRIL 2014, Vol. 136 / 021105-11

Downloaded From: https://tribology.asmedigitalcollection.asme.org on 06/29/2019 Terms of Use: http://www.asme.org/about-asme/terms-of-use

You might also like