You are on page 1of 14

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/258770066

Hydraulic losses of a gearbox: CFD analysis and experiments

Article  in  Tribology International · October 2013


DOI: 10.1016/j.triboint.2013.06.005

CITATIONS READS

26 688

6 authors, including:

Franco Concli Karsten Stahl


Libera Università di Bozen-Bolzano Technische Universität München
119 PUBLICATIONS   264 CITATIONS    216 PUBLICATIONS   254 CITATIONS   

SEE PROFILE SEE PROFILE

Bernd-Robert Hoehn Hansjoerg Schultheiss


Technische Universität München Technische Universität München
155 PUBLICATIONS   652 CITATIONS    12 PUBLICATIONS   67 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Additive Manufacturing in Machine Design View project

Speed2E View project

All content following this page was uploaded by Franco Concli on 07 November 2017.

The user has requested enhancement of the downloaded file.


Hydraulic losses of a gearbox: CFD analysis and experiments
 

Prof. Ing. Carlo Gorla1, Ing. Franco Concli1


Prof. Dr.-Ing. Karsten Stahl2, Prof. i. R. Dr.-Ing. Bernd-Robert Höhn2, Dr.-Ing. Klaus Michaelis2 , Dipl.-Ing. Hansjörg
Schultheiß2,M.Sc. Johann-Paul Stemplinger2
1 Politecnico di Milano, dipartimento di meccanica, via La Masa 1, 20156 Milano, Italy
2 FZG, Gear Research Center, Technical University of Munich, Boltzmannstraße 15, D-85748 Garching bei München, Germany
 

Prof. Ing. Carlo Gorla carlo.gorla@polimi.it +39 02 2399 8223

Ing. Franco Concli franco.concli@mail.polimi.it +39 02 2399 8223

Prof. Dr.-Ing. Karsten Stahl. stahl@fzg.mw.tum.de +49 (089) 289-15805

Prof. i.R. Dr.-Ing. Bernd-Robert Höhn fzg@fzg.mw.tum.de +49 (089) 289-15806

Dr.-Ing. Klaus Michaelis michaelis@fzg.mw.tum.de +49 (089) 289-15809

Dipl.-Ing. Hansjörg Schultheiß schultheiss@fzg.mw.tum.de +49 (089) 289-15837

M.Sc. Johann-Paul Stemplinger stemplinger@fzg.mw.tum.de +49 (089) 289-15822

Abstract: Efficiency is becoming a main concern in the design of power transmissions. It is


therefore important, especially during the design phase, to have appropriate models to predict the
power losses. For this reason CFD (computational fluid dynamics) simulations were performed in
order to understand the influence of geometrical and operating parameters on the losses in power
transmissions. The results of the model were validated with experimental results.

Keywords: Gear, Power losses, efficiency, CFD, lubrication


 

1. INTRODUCTION

Efficiency is becoming more and more a main concern in the design of power transmissions and appropriate
models to predict power losses are fundamental in order to reduce them, starting from the earliest stages of
the design phase. Power losses of gearboxes are generally classified according to [1], taking into account the
machine elements which are responsible for them and their dependency or non-dependency from the load.
Gear power losses are strongly related to lubrication, with those load dependent coming from the frictional
effects in the lubricant film and those load independent mainly deriving from squeezing, churning and
windage effects.
Some models, obtained on the basis of experimental tests, can be found in literature which describe the
influence of gear geometric and kinematic parameters on hydraulic losses like for instance those proposed by
Mauz [2], who has concentrated on hydraulic losses, by Dawson [3] et al. ,who have concentrated on
windage losses or by Seetharaman et al. [4] , who have concentrated on churning losses.
Nevertheless the authors maintain that a deeper understanding of the physical phenomena responsible for
gear losses is still needed in order to improve existing models and CFD simulation can be an effective
approach for such investigation.
Marchesse et al. [5], on the basis of a state of the art on the application of CFD to gear power losses, applied
CFD models to study windage losses of gears and have validated their results by means of experimental
tests.
Hill et al. [6] studied trough CFD simulations the influence of different shrouding configurations on the
windage power losses.
Diab et al. [7] present a number of preliminary experimental and theoretical findings on the prediction of
windage losses. A dimensional analysis has been carried out and a quasi-analytical model considering in
detail the fluid flow on the gear faces and inside the teeth has been developed.

Concli et al. [8 to 13] applied CFD models to study oil squeezing power losses of gears and churning power
losses of planetary speed reducers and have validated their results by means of experimental tests.
A large amount of experimental data on no-load power losses is still available at FZG, resulting from several
years of tests [14], with either single discs, single gears and two meshing gears immersed in oil, and this data
cover the influence of several parameters like outside diameter, face width, helix angle, temperature and oil
type. In order to improve the understanding of the mechanisms involved in hydraulic losses, CFD
simulations have been run in order to investigate the effect of the same parameters and the numerical results
have been compared with the available test results.
In the first phase of the activities, which is presented in [15], single discs and gears have been considered.
The quite good accordance between CFD simulation and experimental tests has confirmed that CFD
represents an effective approach to study windage power losses. In this paper, the effect of the tip diameter,
the face width and the rotational speed on the hydraulic losses (windage + squeezing) on two meshing gears
have been investigated by means of CFD simulations and validated with experimental data. The results
appear in good agreement.

Nomenclature load dependent power losses of gears [W]


-3
gear width [mm] [m ] load independent power losses of gears [W]
-3
  tip diameter [mm] [m ] other generic power losses [W]
  normal module [ ] hydraulic torque [Nm]
  rotational speed [rpm] [2 -1
rad s ]   number of teeth [ ]
power losses [W]   pressure angle [°] [ rad]
load dependent power losses of bearings [W]   helix angle [°] [ rad]
load independent power losses of bearings [W]   dynamic viscosity [Kg m-1 s-1]
power losses of the seals [W]   density [Kg m3]

2. COMPOSITION OF GEARBOX POWER LOSSES


According to [1] the power losses of gears can be subdivided into load dependent and load independent
losses and according to their origins

[1]

In equation [1] the subscripts , , and are related to gears, to bearings, to contact seals and to other
factors respectively. The subscript indicates load independent power losses.
are the load dependent power losses of gears and arise primarily from the sliding between the flanks of
the gear teeth.
are the load dependent power losses of bearings and are also related to sliding between the rolling
elements and the rings.
are the load independent power losses of bearings and are, among others, related to viscous effects due
to the lubrication.
are the power losses of the seals and are related to sliding.
are other generic power losses.
are the load independent power losses of gears due to the interaction between the lubricant and the
rotating/moving elements. These losses are the sum of squeezing, churning and windage power losses. The
lubricant squeezing power losses are related to the fact that the gap at the mesh position is changing its
volume during the engagement causing an overpressure that squeezes the oil primarily in the axial direction
[8]. The churning and windage power losses are related to the viscous and pressure effects of the lubricant on
the moving/rotating elements [9, 11]. For dip-lubricated power transmissions this kind of losses are an
important part of the total losses. For this reason the authors have studied this kind of losses for a simple
geometry and under different operating conditions.
3. PROBLEM DESCRIPTION

The analyzed gearbox is designed for a high pressure multiphase pump. This kind of screw pumps are
developed in order to operate on the bottom of the sea up to 3000 m depth, pumping gas, water and oil-
mixtures. The gearboxes operating in such environments are typically completely filled with lubricant. This
fact has two implications: while the oil allows to compensate the external pressure of the sea water avoiding
the collapse of the case, on the other side its quantity induces significant windage losses. In this case the
speed reducer is made by a single stage and operates with circumferential velocities up to 38 m/s.

3.1 Experimental tests


3.1.1 Geometry

Figure 1a shows the layout of the test rig.

a) b)

Figure 1: a) layout of the test rig, b) picture of the gearbox with 5 gear pairs mounted

The gearbox is connected to an electric motor. By means of a torque meter it is possible to measure the
resistant torque. The gearbox is equipped by a manometer that allows to measure the internal pressure of the
lubricant. Inside the gearbox, the two tested gears are mounted on two cantilever  shafts supported by
bearings. A bulkhead allows to change the depth of the case. In previous steps of the research, the tests have
been performed with only 1 gear [11]. In this phase of the research, both the experimental tests and the
numerical simulation have been performed with a gear pair as shown in figure 1b. The geometrical
properties of the adopted gears are summarized in the table 1.

Table 1: geometrical properties of the analyzed gears


Chordal thickness
Number of teeth

Normal module
Pressure angle
[mm]

Centre distance

Root diameter
Tip diameter

[mm]
[mm]

Helix angle

[mm]
[mm]
[mm]
[°]

[]
[°]
Width

Reference case 40 102.5 0 20 23 4 91.5 91.5 5.35


Tip diameter influence 40 96.5 0 20 23 4 91.5 91.5 5.35
Tooth width 20 102.5 0 20 23 4 91.5 91.5 5.35
 

A cooling/heating system was also present in order to impose and maintain a constant temperature during the
tests.
3.2 Numerical model
3.2.1 Geometry

For the numerical simulations some geometry simplifications have been adopted (some minor detail like
chamfers and bearings have not been completely modeled). Despite of the cantilever shafts, it was chosen to
consider the model as symmetric. This simplification was adopted in order to reduce the volume of the
considered domain, with the reduction of the amount of cells needed for the numerical discretization.

Furthermore  the  gears  have  been  moved  a  little  apart  so they don’t actually  touch  in  the  CFD  model  in 
order to guarantee the continuity of the computational domain. This is necessary in order to guarantee the 
continuity of the mesh. 

After this simplifications, the virtual model appears as shown in figure 2.

Figure 2: geometry of the numerical model

3.2.1 Mesh

In order to discretize the numerical model, two kinds of mesh have been used. Near the walls of the gearbox
case, a tetrahedral mesh was used. For complex geometries, in fact, this mesh topology can often be created
with far fewer cells than the equivalent mesh consisting of quadrilateral/hexahedral elements. This is because
the tetrahedral mesh allows clustering of cells in selected regions of the flow domain. Structured hexahedral
meshes will generally force cells to be placed in regions where they are not needed.

In the region of the meshing and in proximity of the gears (in the inner partition visible in figure 2), a
triangular extruded mesh was used. The choice of this kind of mesh was mandatory in order to be able,
during the calculation, to adapt the mesh to the geometrical changes due to the rotation of the gears. This
mesh topology, called swept mesh, consist in creating a 2D triangular mesh on one side of the region, known
as the source side, and then copying the nodes of that mesh, one element layer at a time, until the final side,
known as the target side, is reached. Due to the form of the elements on the source side (triangles), it allows
to create meshes with low skewness also for complex geometries like in this case.

For the analyzed geometry, 6.5x106 elements have been used.

In order to update the volume mesh in the deforming regions subject to the motion defined at the boundaries
(the gear surfaces), a spring-based smoothing method and a 2.5D remeshing technique were both adopted. In
the Ansys spring-based smoothing method, the edges between any two mesh nodes are idealized as a
network of interconnected springs. The initial spacing of the edges before any boundary motion constitute
the equilibrium state of the mesh. A displacement at a given boundary node will generate a force
proportional to the displacement along all the springs connected to the node. Using Hook's Law, the force on
a mesh node can be written as

⃗ ∑ ( ⃗ ⃗) [2]

where ⃗ and ⃗ are the displacements of node and node , is the number of neighboring nodes
connected to node , and is the spring constant (or stiffness) between node and its neighbor . The
spring constant for the edge connecting nodes and is defined as

[3]
√| ⃗ ⃗ |

At equilibrium, the net force on a node due to all the springs connected to the node must be zero. This
condition results in an iterative equation such that

∑ ⃗
⃗ [4]

Since displacements are known on the boundaries (after boundary node positions have been updated), this
last equation is solved using a Jacobi sweep on all interior nodes. At convergence, the positions are updated
such that

⃗ ⃗ ⃗ [5]

where and are used to denote the positions at the next time step and the current time step,
respectively.

On zones with a triangular or tetrahedral mesh, the spring-based smoothing method is normally used. When
the boundary displacement is large compared to the local cell size, the cell quality can deteriorate or the cells
can become degenerate. This will invalidate the mesh and consequently, will lead to convergence problems
when the solution is updated to the next time step. To circumvent this problem, it is possible to locally
remesh some cells agglomerate that did not satisfy the skewness criteria or the maximum and minimum
length allowed for the cells. In order to do that a 2.5D remeshing method has been adopted.

Figure 3 shows some details of the mesh.


Figure 3: details of the mesh

3.4 Solver settings


3.4.1 Averaged Navier Stokes Equations
The CFD is based on some differential equations. For a generic element it is possible to write five equations.
The first equation is the averaged mass conservation equation for no-stationary incompressible flows which
can be written as follows
〈 〉
[6]

where is the Cartesian coordinate and is the velocity component. The second equation is the averaged
momentum conservation equation and can be written as follows

〈 〉 〈 〉
( 〈 〉) ( 〈 〉〈 〉) 〈 〉 [ ( ) ] [7]

Where is the pressure and are the Cartesian coordinates and is the density. The additional term
that compare in the averaged equation (in comparison to the non-averaged transport equations) is called
unresolved term or Reynolds term. The averaging process produces a set of equations that is not closed. For
this reason a turbulence model is needed in order to be able to solve the system of equations. The unresolved
or Reynolds terms are expressed by using the eddy-viscosity hypothesis as

〈 〉 〈 〉
〈 〉 ( ) [8]

where is the turbulent kinetic energy and the eddy viscosity. The energy equation was not activated in
the given model: the operating temperature was defined a priori and consequently the properties of the fluid
do not change during the calculations. For the pressure-velocity-couplig a SIMPLE scheme was adopted as
suggested for flows in closed domains [12]. This algorithm uses a relationship between velocity and pressure
corrections to enforce mass conservation and to obtain the pressure field. For incompressible flows, the
solution of the system of equations is accoomplished by the fact that there are not equations where the
pressure is explicity defined. To calcluate it, the continuity equation is substituted with an equation for the
pressure; with some manipulation the pressure appears like unknown term in the momentum equation.

3.4.2 Turbulence models


Turbulent flows are characterized by fluctuating velocity fields. These fluctuations mix transported quantities
such as momentum, energy, and species concentration, cause the transported quantities to fluctuate as well.
Since these fluctuations can be of small scale and high frequency, they are too expensive from the
computational point of view to be simulated directly in practical engineering calculations. Instead, the
instantaneous exact governing equations can be averaged to remove the small scales, resulting in a modified
set of equations that are computationally less expensive to solve. However, the modified equations contain
additional unknown variables, and turbulence models are needed to determine these variables in terms of
known quantities. The simplest "complete models'' of turbulence are two-equation models in which the
solution of two separate transport equations allows the turbulent velocity and length scales to be determined
independently. This model is a semi-empirical model based on model transport equations for the turbulence
kinetic energy ( ) and its dissipation rate ( ) [13]. The RNG - model assumes that the eddy viscosity is
related to the turbulence kinetic energy and dissipation via the relation

[9]

The model transport equation for is derived from the exact equation, while the model transport equation for
was obtained using physical considerations and bears little resemblance to its mathematically exact
counterpart [14]. In the derivation of the - model, the assumption is that the flow is fully turbulent, and
the effects of molecular viscosity are negligible.

( ) ( ) [ ] [10]

( ) ( ) [ ] ( ) [11]

In these equations, represents the generation of turbulence kinetic energy due to the mean velocity
gradients. is the generation of turbulence kinetic energy due to buoyancy effect. represents the
contribution of the fluctuating dilatation in compressible turbulence to the overall dissipation rate. , ,
and are constants. and are the turbulent Prandtl numbers for and , respectively. The quantities
and are the inverse effective Prandtl numbers for and , respectively. and are user-defined
source terms. is the additional term that characterizes the RNG-k-ε-model. is derived from a rigorous
statistical technique called renormalization group theory and improves the accuracy of the Standard-k-ε-
model. Some other more complex turbulent models exist, but some authors claim that this model offers
improved accuracy in rotating flows [14]. For this reason this turbulence model was selected to perform the
simulations.

3.4.3 Spatial discretization


By default, the solver stores discrete values of the scalar at the cell centers. However, face values are
required for the convection terms and must be interpolated from the cell center values. This is accomplished
using an upwind scheme. In order to solve the differential equations, the authors have chosen to use second
order upwind schemes. For the determination of the gradient, a least squares cell based evaluation was used.
In this method the solution is assumed to vary linearly between two cell centroids.

3.4.4 Boundary conditions


All the boundaries were set to a no slip condition. To describe properly the velocity profile in the normal
direction, an enhanced wall treatment was used. This technique calculates the velocity at the elements’ center
and then reconstructs the velocity profile starting from these quantities. The y+ value for this simulations was
approximately equal to one.

3.4.5 Time steps


The determination of the time step size is based on the estimation of the truncation error associated with the
time integration scheme. If the truncation error is smaller than a specified tolerance, the size of the time step
is increased; if the truncation error is greater, the time step size is decreased [15].
3.4.6 Torque calculations
The total moment vector about a specified center is computed by summing the cross products of the
pressure and viscous force vectors for each face with the moment vector ⃗⃗ , which is the vector from the
specified moment center to the force origins

⃗⃗⃗⃗ ⃗⃗ ⃗⃗ ⃗⃗ ⃗⃗ [12]

where ⃗⃗ and ⃗⃗ represent the pressure force vector and the viscous force vector respectively.

4. OPERATING CONDITIONS
With CFD analysis the effect of different parameters was investigated. Table 1 summarizes the combination
of parameters adopted in the different simulations.

Table 1: combination of parameters adopted for the CFD simulations

Number of teeth

Normal module

Rotational speed
Pressure angle
[mm]

Tip diameter

Oil dynamic
[mm]

Oil density
[mm]

[Kg/m s]
Helix angle

[Kg/m³]

viscosity
[°]

[]

n[rpm]
[°]
Width

Reference case 1 40 102.5 0 20 23 4 2000 824.5 5.463e-3


Reference case 2 40 102.5 0 20 23 4 5000 824.5 5.463e-3
Reference case 3 40 102.5 0 20 23 4 8000 824.5 5.463e-3
Tip diameter influence 1 40 96,5 0 20 23 4 2000 824.5 5.463e-3
Tip diameter influence 2 40 96,5 0 20 23 4 5000 824.5 5.463e-3
Tip diameter influence 3 40 96,5 0 20 23 4 8000 824.5 5.463e-3
Tooth width 1 20 102.5 0 20 23 4 2000 824.5 5.463e-3
Tooth width 2 20 102.5 0 20 23 4 5000 824.5 5.463e-3
Tooth width 3 20 102.5 0 20 23 4 8000 824.5 5.463e-3

The real gearbox used for the experiments was developed to operate on the bottom of the sea and an internal
pressure (6 bar) was needed in order to compensate the external water pressure at operating depth. The
experiments have been therefore performed with the operating conditions of 6 bar. The pressurization of the
gearbox was made after some preliminary experiments without pressure. The choice to operate with 6 bar
was taken in order to be sure to have only oil in the gearbox avoiding the presence of air bubbles. Moreover,
the level of the static pressure, does not affect the results.

Oils with different grades of viscosity have at this point of the research not been tested because the influence
of the viscosity is negligible as reported by the authors in [11].

5. EXPERIMENTAL TESTS
In order to validate the results of the CFD simulations, some experiments were performed. On the driving
shaft, an own developed torque meter was mounted. This instrument measures the resistant torque and the
rotational velocity so it is possible to evaluate the power losses. The torque meter was equipped with a DMS
Wheatstone bridge that was calibrated between 13,23 – 200,12 Nm.

Due to the configuration of the system, it is not possible to measure directly the load independent power
losses due to the presence of the losses of the bearings. For this reason some additional tests were performed
without the gears and with air instead of oil in the gearbox so as to be able to measure the power losses
generated by the bearings.
The bearings are however almost completely not loaded so their losses are of a lower order of magnitude
with comparison to the load-independent losses of gears. Making the difference between the total losses and
the losses of the bearings, it is possible to evaluate the losses of the gears (windage + squeezing).

The oil temperature during the test was constantly monitored with a temperature sensor in the oil sump. The
test gear box had integrated heating as well as cooling elements that were automatically activated when the
measured oil temperature deviated from the desired value by more than ± 3 .

6. RESULTS

From figure 4 it is possible to have an idea of the fluxes of lubricant inside the gearbox. An ideal particle of
fluid escapes the gear radially and then, after a loop, it comes again in contact with the gear from the axial
direction. Due to the rotation of the gear, there are no zones in the gearbox where the lubricant is steady but
the complete domain is involved in the lubricant flux.

Figure 4: velocity streamlines; b=40mm; d=102.5mm; T=90°C, n=8000rpm

250 120
Experimental Experimental
Hydraulic torqueTVZ0,H [Nm]

Hydraulic torque TVZ0,H [Nm]

Polynomial extrapolation Polynomial extrapolation


200 CFD 100 CFD
da=102.5mm; da=96.5mm;
b=40mm; 80 b=40mm;
150 100% Oil 100% Oil

60
100
40

50 20

0 0
0 1500 3000 4500 6000 7500 9000 0 1500 3000 4500 6000 7500 9000
Rotational speed n [rpm] a)  Rotational speed n [rpm] b)
120
Experimental
Hydraulic torqueTVZ0,H [Nm]

Polynomial extrapolation
100 CFD
da=102.5mm;
80 b=20mm;
100% Oil

60

40

20

0
0 1500 3000 4500 6000 7500 9000
Rotational speed n [rpm]  c)

Figure 5: Comparison between the numerical and the experimental results*


* (simulations have been performed on a 16 cpus workstation; the computational time was approximately of 20 hours for each simulation)

Figure 5 shows the comparison between the numerical and the experimental results in terms of no load loss
torque (given by windage and squeezing effects) versus rotational speed . Figure 5a represent the
“standard” case with the tip diameter and the tooth width . The continuous line
represents the results of the experiments without the losses of the bearings and the striped lines a polynomial
extrapolation of the experimental results. The experimental results that appear in figure 5 are obtained as
difference between the total measured losses in the working configuration (including therefore the bearing
losses) and the losses measured with only the shafts (no gears) and air instead of oil in the gearbox. In this
manner it is possible to separate the power losses due to the rotation of the gears in the oil from the losses
caused by the bearings. It is however necessary to specify that in this second measurement, the absence of
the lubricant may minimally affect the losses caused by the bearings: the inner race of the bearings, in fact,
during the rotation is laterally in contact with air instead of oil and this leads to a minimal underestimation of
the losses of the bearing. The rectangles represent the results of the numerical simulations. It can be seen that
the CFD results are quite in good agreement with the experimental ones. As expected, the losses evolve with
angular velocity power three.

Experimental da=102.5mm Experimental b=40mm


Polynomial extrapolation da=102.5mm 300 Polynomial extrapolation b=40mm
300 CFD da=102.5mm CFD b=40mm
Hydraulic losses TVZ0,H [Nm]
Hydraulic losses TVZ0,H [Nm]

Experimental da=96.5mm Experimental b=20mm


250 Polynomial extrapolation da=96.5mm 250 Polynomial extrapolation b=20mm
CFD da=96.5mm CFD b=20mm

200 b=40mm; 200 da=102,5mm;


100% Oil 100% Oil

150 150

100 100

50 50

0 0
0 1500 3000 4500 6000 7500 9000 0 1500 3000 4500 6000 7500 9000
Rotational speed n [rpm] a) Rotational speed n [rpm] b)

Figure 6: a) Effect of the tip diameter on the power losses, b) effect of the tooth width on the power losses
Figure 6a shows the effect of the tip diameter on the no load torque. It can be seen that the tip
diameter has a big influence on the losses: a decrease of the diameter from 102.5mm to 96.5mm causes a
decrease of the losses of about 53%. As expected the windage losses evolve with tangential speed power
three. For this reason, the tip diameter plays an important role (together with the rotational speed): increasing
the tip diameter means increasing the mean relative velocity between fluid and gear and therefore the losses.

Figure 6b shows the influence of the tooth width on the power losses. A halving of the tooth width causes an
decrease of about 54% of the losses. The losses would not be halved even if the eternal radial surface (tooth
surface) of the gears is halved due to the side effects that remain almost the same (interaction between the
side of the gears and the lubricant).  

300 300
CFD da=102.5mm CFD b=40mm
CFD da=96.5mm CFD b=20mm

Resistant torque TVZ0 [Nm]


Resistant torque TVZ0 [Nm]

250 250 CFD 1GEAR b=40mm


CFD 1GEAR da=102.5mm
CFD 1GEAR da=96.5mm CFD 1GEAR b=20mm
200 200 da=102,5mm;
b=40mm;
100% Oil 100% Oil
150 150

100 100

50 50

0 0
0 1500 3000 4500 6000 7500 9000 0 1500 3000 4500 6000 7500 9000
Rotational speed n [rpm] Rotational speed n [rpm]
a) b)

Figure 7: Comparison of the hydraulic losses generated by 2 engaging gears and a single rotating gear: a)
effect of the tip diameter on the power losses, b) effect of the tooth width on the power losses

100 100
CFD da=102.5mm CFD b=40mm
Squeezing losses TVZ0,S [Nm]
Squeezing losses TVZ0,S [Nm]

CFD da=96.5mm CFD b=20mm


80 80
b=40mm; da=102,5mm;
100% Oil 100% Oil
60 60

40 40

20 20

0 0
0 1500 3000 4500 6000 7500 9000 0 1500 3000 4500 6000 7500 9000
Rotational speed n [rpm] a) Rotational speed n [rpm] b)

Figure 8: Comparison of the squeezing losses generated by 2 engaging gears and a single rotating gear: a)
effect of the tip diameter on the power losses, b) effect of the tooth width on the power losses (the
squeezing losses have been calculated as the difference between the hydraulic losses generated by 2
engaging gears and the double of the windage losses generated by a single rotating gear.)

From figure 7 it appears, as expected, that the hydraulic losses generated by two meshing gears are more
than the double of the windage losses generated by a single rotating gear. This fact is due to the additional
squeezing losses (figure 8) that arise when a second gear is present. During the engagement, in fact, the
volume of the gap between the teeth reduces and expands. This implies a first overpressure in the gap that
squeezed out the lubricant and a second reduction of the pressure in the gap so that a fluid flow takes place
from the oil bath to the gap. Due to the viscous effects of the lubricant, this phenomena induces additional
losses.

7. CONCLUSIONS
The results of the experiments confirm that the CFD represent a valid method to predict power losses. The
error in the predictions for the analyzed cases is always lower than 8%. This little difference can be
explained with the fact that the geometry has been simplified. The simulations were performed for different
geometries; in particular the tip diameter has been changed from 96.5 to 102.5mm while the tooth width has
been changed from 20 to 40mm. Also the rotational speed of the gears has been varied from 0 up to
8000rpm.

This three analyzed parameters have different influences on the losses. The losses increase with the velocity
with power three. The losses appear to be related with the tooth width in a linear manner. Finally the losses
appear to increase significantly with the tip diameter that is, together with the rotational speed, the most
influencing parameter.

Since CFD proves to be a valid tool to predict the load independent losses, it is planned to investigate the
influence of other geometrical and operating parameters as well as other sources of losses like the churning
ones.

BIBLIOGRAPHY
[1] Niemann, G., Winter, H.: Maschinenelemente Band II – Getriebe allgemein, Zahnradgetriebe –
Grundlagen, Stirnradgetriebe, Springer-Verlag, Berlin Heidelberg Ney York Tokio 1983

[2] Mauz, W.: Hydraulische Verluste von Stirnradgetrieben bei Umfangsgeschwindigkeiten bis 60 m/s.
Bericht des Institutes für Maschinenkonstruktion und Getriebebau Nr. 159, Universität Stuttgart 1987.

[3] Dawson, P.H., Windage loss in larger high-speed gears, Proc. Inst. Mech. Eng., 198A (1), pp.51-59.

[4] Seetharaman S., Kahraman A. Moorhead M. D., Petry-Johnson T. T., Oil Churning Power Losses of a
Gear Pair: Experiments and Model Validation, Journal of Tribology Volume 131, Issue 2, April 2009, Pages
022202-1 - 022202-10

[5] Marchesse Y., Changenet C., Ville F., Velex P., Investigation on CFD simulation for predicting windage
power losses in spur gears, ASME Journal of Mechanical Design, 2011, 133(2), 7 pages.

[6] Hill, M.J.a , Kunz, R.F.b , Medvitz, R.B.b , Handschuh, R.F.c , Long, L.N.a , Noack, R.W.b , Morris,
P.J.a , CFD analysis of gear windage losses: Validation and parametric aerodynamic studies, Journal of
Fluids Engineering, Transactions of the ASME Volume 133, Issue 3, 2011, Article number031103

[7] Diab, Y.a, Ville, F.a, Velex, P.a, Changenet, C.b, Windage losses in high speed gears-preliminary
experimental and theoretical results, Journal of Mechanical Design, Transactions of the ASME Volume 126,
Issue 5, September 2004, Pages 903-908

[8] Concli, F., Gorla, C.: Oil squeezing power losses of a gear pair: a CFD analysis, 9th International
Conference on Advances in Fluid Mechanics - Advances in Fluid Mechanics IX, WIT Transactions on
Engineering Sciences Volume 74, 2012, Pages 37-48, WIT 2012 (ISSN: 17433533 ISBN: 978-184564600-4)

[9] Concli, F., Gorla, C.: Computational and experimental analysis of the churning power losses in an
industrial planetary speed reducers, 9th International Conference on Advances in Fluid Mechanics -
Advances in Fluid Mechanics IX, WIT Transactions on Engineering Sciences Volume 74, 2012, Pages 287-
298, WIT 2012 (ISSN: 17433533 ISBN: 978-184564600-4)

[10] Höhn B.-R., Michaelis K., Otto H.-P.: Influence on no-load gear losses, Ecotrib 2011 Conference,
Proceedings Vol.2 pp.639-644, 2011

[11] Gorla C., Concli F., Stahl K., Höhn B.-R., Michaelis K., Schultheiß H., Stemplinger J.-P., CFD
simulations of splash losses of a gearbox, “Tribological Challenges in Mechanical Transmissions” special
issue of “Advances in Tribology”, Article number 616923, HINDAWI 2012 (ISSN: 1687-5915)

[12] Versteeg, H.K., Malalasekera, W.: An introduction to computational fluid dynamics – The finite volume
method, Longman Group, London 1995method, Longman Group, London 1995

[13] Launder B. E., Spalding D. B., The Numerical Computation of Turbulent Flows, Computer Methods in
Applied Mechanics and Engineering, 1974, 3, p. 269–289.

[14] Yakhot, V., Orszag, S.A., Thangam, S., Gatski, Development of turbulence models for shear flows by a
double expansion technique, Physics of Fluids A, Vol. 4, No. 7, pp1510-1520.

[15] Gresho P. M., Lee R. L.,Sani R. L. On the Time-Dependent Solution of the Incompressible Navier-
Stokes Equations in Two and Three Dimensions, Recent Advances in Numerical Methods in Fluids.
Pineridge Press, Swansea, U. K.. 1980.

View publication stats

You might also like