Вы находитесь на странице: 1из 61

LULEL

2003:15
UNIVERSITY
OF TECHNOLOGY

Performance Analysis of a Low-Speed High-Torque


Hydrostatic Transmission Unit

Håkan Olsson

Department of Applied Physics and Mechanical Engineering


Division of Machine Elements

2003:15 • ISSN: 1402 - 1757 • ISRN: LTU - LIC - - 03/15 - - SE


Performance Analysis of a Low-Speed High-Torque
Hydrostatic Transmission Unit

Håkan Olsson

ITH, The Institute for Applied Hydraulics


Örnsköldsvik, Sweden
and •
Department of Mechanical Engineering
Division of Machine Elements
Luleå University of Technology
Luleå, Sweden

Örnsköldsvik and Luleå, February 2003


to Klara
ABSTRACT

This thesis concerns a study of the performance of an industrial low-speed high-torque hydrostatic
drive system. This type of hydrostatic transmission is commonly used in continuous operation in a
wide range of heavy-duty drive applications. In many applications the transmissions have to
compete with e.g. electromechanical drives, such as DC or AC electric motors combined with
gearboxes. In such situations, energy efficiency is a key selection criterion in that even a small
increase in the efficiency of high power industrial drives would give substantial savings. Apart from
efficiency, lifetime and reliability requirements are important parameters for industrial drive
systems, as unplanned stops in industrial working processes can be very costly.

The work presented in this thesis is primarily focused on analysing the efficiency behaviour in the
transmission, both on the system level and on the component level. Attention has also been paid to
lifetime issues, with special emphasis on wear occurring in a sliding contact in a radial piston
hydraulic motor.

In Paper A the distribution of power losses in a variable axial piston swash plate pump is
investigated. The pump under study is commonly used in stationary industrial hydrostatic
transmission systems. The churning losses in the pump have been estimated experimentally by
measurements in a test rig. The leakage flow and the power losses in the contacts between the
piston and the cylinder and between the slipper and the swash plate respectively were simulated
with the help of the simulation tool CASPAR. For the pump studied, the churning losses are
significant under the operating conditions typically occurring in industrial drive applications. The
simulation results indicate that the leakage to the pump casing mainly originates from the gaps
between the pistons and their respective cylinders and between the slippers and swash plate.

The aim of Paper B is to study two sliding contacts inside a radial piston hydraulic motor and
investigate their influence on the torque and power losses. Moreover, it is investigated whether and
when a change in the lubrication regime can be expected in these contacts. This is accomplished by
a combination of experimental and theoretical studies, with a special focus on two lubricated sliding
contacts: the distributor valve contact and the piston/cam roller contact. The theoretical analysis of
the contacts indicates, among other things, that the piston/cam roller contact can enter the mixed
lubrication regime at low motor speeds.

At low running speeds, an increased wear rate has been noted in the contact between the cylinder
bore and the piston skirt in a radial piston hydraulic motor. Paper C describes a comparative
investigation into different hydraulic fluids' friction properties and wear protection abilities. To
simulate the contact between the cylinder bore and the piston skirt in the hydraulic motor, tests were
performed in a reciprocating test rig where the contact geometry was of the cylinder-plate type. In
the model test a synthetic ester aimed at meeting the conditions in water turbine applications
received a top ranking regarding both friction and wear protection properties.
CONTENTS

1 INTRODUCTION 1
1.1 BACKGROUND 1
1.2 OBJECTIVES 2

2 EFFICIENCY 2

2.1 EFFICIENCY IN DISPLACEMENT MACHINES 2


2.2 EFFICIENCY IN HYDROSTATIC TRANSMISSIONS 3
2.2.1 Closed hydrostatic transmissions 3
2.2.2 Efficiency models 6
2.2.3 Simulation tools 7
2.2.4 Experimental steady state measurements 8
2.2.5 Hydraulic fluids 8
3 LUBRICATION AND WEAR 11
3.1 LUBRICATION REGIMES 11
3.2 WEAR MECHANISMS 12
3.3 WEAR TESTS FOR HYDRAULIC FLUIDS 12

4 SUMMARY OF PAPERS 13

5 CONCLUDING REMARKS 14

6 REFERENCES 15

PAPER A
PAPER B
PAPER C
APPENDED PAPERS

A. Olsson, H., Power Losses in an Axial Piston Pump Used in Industrial Hydrostatic
Transmissions, submitted for presentation at the 8th Scandinavian International Conference on
Fluid Power, SICFP'03, Tampere, Finland, May 7-9, 2003.

B. Dahldn, L., Olsson, H., Analysis of Two Sliding Contacts inside a Radial Piston Hydraulic
Motor, The 5th JFPS International Symposium on Fluid Power, Nara, Japan, 2002.

C. Olsson, H., Ukonsaari, J., Wear Testing and Specification of Hydraulic Fluid in Industrial
Applications, The 10th Nordic Symposium on Tribology, Stockholm, Sweden, June 9-12,
2002.
1 INTRODUCTION
Hydraulics, often referred to as "Fluid Power", is a branch of technology where a confined liquid is
used for transmitting energy, i.e. force and motion. The term originates from the Greek words for
water (hydro) and pipe (aulos), and hydraulics concerns the application of the laws of
hydromechanics. Hydromechanics can be divided into two areas: hydrodynamics and hydrostatics.
Hydrodynamics is based on liquids in motion, with the energy conversion being accomplished by
dynamic machines (i.e. centrifugal pumps and turbines). Hydrostatics deals with liquids under
pressure, with displacement machines being used for the transformation of energy. When referring
to hydraulics (or fluid power) in everyday speech, one often means hydrostatics.

The main prerequisite for the successful use of hydrostatic systems is the use of displacement
machines. Displacement machines can be of different types and are mainly classified into two
groups. This classification is based on the type of movement of the displacement element that
realizes the increase and decrease of the displacement chamber [1].
• Machines whose displacement element uses rotary motion. The element can be in the form
of teeth, screws or vanes and the most common machines in this group are thus gear, screw
and vane machines. Vane machines can be made either with a variable or a fixed
displacement, i.e. with or without the possibility of varying the geometrical displacement
volume. The other two machine types in this group only have a fixed displacement volume.
• Machines whose displacement element uses reciprocating motion. Here the element is
mostly a piston and this group of displacement machines is therefore also known as piston
machines. The most common machines of this type are axial piston machines and radial
piston machines. They can be made both with a fixed and with a variable displacement
volume.

The most commonly used displacement machines, especially in the region of higher operating
pressure (above 20 MPa), are piston machines. They possess many advantages, for instance high
power densities and favourable efficiencies. In the case of axial piston machines it is also easy to
realize effective measures to vary the displacement volume. Axial piston machines are normally
either of a swash plate design or of a bent axis design. Radial piston machines can typically be of
the eccentric type or of the cam lobe design. For a deeper insight into the characteristics of the
different displacement machines, see [1].

1.1 Background

Hydrostatic transmissions are hydraulic drives where positive displacement hydraulic pumps and
motors transfer rotary power by means of fluid under pressure. Hydrostatic transmissions are
utilized both in mobile and in industrial drive applications, and they can typically be used in heavy-
duty drives such as logging machines, forest harvesters, excavators, cranes, tunnel-boring machines
etc. In the hydrostatic transmission system under study in this thesis, radial piston motors of the
cam ring type are used as driving units. The hydraulic motors are designed for continuous operation
in low-speed high-torque drive applications where the motors are mounted directly on the shaft of
the machine that is to be driven. The hydraulic motor has a fixed displacement and is connected to
an axial piston swash plate pump with variable displacement in a closed loop system. By varying
the displacement setting of the pump, the speed of the hydraulic motor can be varied. The pump in
these transmissions is normally driven at a more or less fixed speed by an induction electric motor.

Industrial low-speed high-torque hydrostatic drive systems are commonly used in continuous
operation in a wide range of heavy-duty drive applications. Typical examples of industrial areas that
use this type of hydrostatic transmission are cement, pulp and paper, mining and chemical
industries. In many applications within these sectors, hydrostatic transmissions have to compete
with e.g. electromechanical drives, such as DC or AC electric motors combined with gearboxes. In
such situations, energy efficiency is a key selection criterion in that even a small increase in the
efficiency of high power industrial drives would give substantial savings.

The efficiency of a machine or a system of machines concerns the relationship between the ideal,
loss-less working process and the working process including losses. The efficiency of a hydrostatic
transmission is dependent upon the loss characteristics of the individual components. The loss
behaviour in hydraulic machines and systems can be analysed by experimental steady-state
measurements. These measurements provide the basis for the development of theoretical loss
models that can be utilized to study the loss performance of hydraulic units. The advances in the
performance of computers have led to the development of different simulation tools that are used to
estimate individual loss sources in hydraulic machines.

Apart from efficiency, lifetime and reliability requirements are important parameters for industrial
drive systems. Industrial processes are often continuous and unplanned stops in the working process
can be very costly. Consequently, it is important that the drive systems should be robust and reliable
and give the opportunity to take measures for maintenance in a planned way. Hydraulic fluid is used
to transfer energy between the different components in a hydraulic system. Apart from having an
effect on the efficiency performance of the components, the fluid also serves as a lubricant for the
contacts moving relative to each other inside hydraulic components. Hydraulic fluid has therefore
an influence on the amount of wear occurring in hydraulic components. Wear affects the lifetime of
a component and it can also cause a change in its function due to e.g. increase in clearances. Often
the leakage from a hydraulic component increases with time due to wear.

1.2 Objectives
The objective of the present work was to study the performance of a hydrostatic transmission
system used in industrial heavy-duty drive applications. This has been accomplished by using both
practical experiments and theoretical analyses. The work is primarily focused on analysing the
efficiency behaviour in the transmission, both on the system level and on the component level.
Attention has also been paid to lifetime issues, with special emphasis on wear occurring in the
piston cylinder contact in a hydraulic motor.

2 EFFICIENCY
The term "efficiency" is an important parameter of power transmitting machines. Efficiency
concerns the relationship between the ideal, loss-less working process and the working process
including losses.

2.1 Efficiency in displacement machines


A hydraulic displacement machine converts mechanical power into hydraulic power or hydraulic
power into mechanical power, and the total efficiency of a hydraulic displacement machine is
defined as the ratio between these two. Hence, the total efficiency of a hydraulic displacement
pump is defined as the ratio between the hydraulic output power and the mechanical input power.
For a hydraulic displacement motor the ratio is the opposite.

2
The total (overall) efficiency of a displacement machine can be divided into two parts: volumetric
efficiency and hydromechanical efficiency. Volumetric efficiency concerns losses due to leakage
through the different gaps of the displacement machine. Losses resulting from the compressibility
of the fluid also belong to the volumetric losses. Hydromechanical efficiency concerns torque losses
due to different types of friction occurring between the moving parts inside the machine. It also
concerns losses due to fluid flow, i.e. hydrodynamic losses. A more thorough description of
efficiency in displacement machines can be found in e.g. [1].

2.2 Efficiency in hydrostatic transmissions

A transmission is a device that is capable of matching the input torque and speed to the torque and
speed requirements of the output member that drives the load. This can be accomplished in different
ways based on the physical principal used.
• Mechanical transmissions use contact forces or friction forces between mechanical
components.
• Electrical transmissions utilize electromagnetism.
• Fluid transmissions use forces of inertia, viscous forces or hydrostatic pressures.

Closed hydraulic systems with displacement-controlled hydraulic pumps and motors are often
called hydrostatic transmissions.

2.2.1 Closed hydrostatic transmissions

The principal structure of a hydrostatic transmission with angular energy converters is shown in
Fig. 1. In the most general case, the main components of the transmission consist of a variable
displacement hydraulic pump (a) and a variable displacement hydraulic motor (b). To make the
hydrostatic system work properly, the closed main circuit must be supplemented with a number of
auxiliary functions. A charge circuit with a flushing block is required, partly to secure a certain low-
pressure level to avoid the risk of cavitation and partly to obtain an exchange of the fluid in the
closed circuit for cooling and filtration. The charge circuit consists of a charge pump (c), two check
valves (d), a pressure relief valve (e) and a flushing valve (f). To limit the maximum pressure level
in the system, pressure relief valves are connected in parallel between the low- and high-pressure
sides of the transmission (g). This achieves an overload protection.

Driving Driven
machine machine

Fig. 1. Principal sketch of a closed hydrostatic transmission.


3
Some advantages of this system concept compared to a mechanical transmission are:
• a stepless variation of the gear ratio in a large speed range
• easy regulation of the mechanical output variables, torque and speed
• a flexibility which allows the convenient location of the prime mover, load and
interconnections
• overload protection built into the system
• insensitivity to shock loads
• high power-to-weight ratio
• low inertia of the rotating parts, which allows rapid starting, stopping and reversing.

Hydrostatic transmissions also offer the possibility of energy storage. An example of this can be
found in Rydberg [2], who has studied the possibilities of reducing the fuel consumption in city
buses by storing kinematic energy normally dissipated as brake heat and re-using it for acceleration.
A regenerative hybrid drive system is used which includes an energy storage device and a
continuously variable hydrostatic transmission. The drawbacks of hydrostatic systems are mainly
related to price and the fact that a hydrostatic transmission is often less efficient than, for instance, a
mechanical transmission.

The total efficiency of a hydrostatic transmission is determined by the system solution and the
efficiency of each individual component in the transmission. It is also Fig. 2 shows results from
laboratory measurements of the efficiency of an electrically driven hydrostatic transmission [3]. The
transmission consists of an induction electric motor, a variable displacement axial piston pump of
the swash plate type and a radial piston hydraulic motor of the cam ring type. The hydraulic pump
runs at a constant speed (1500 rpm) and is equipped with external charge and pilot pumps mounted
in tandem to the main pump. In Fig. 2 the total efficiency of the transmission is shown, together
with the total efficiency of the three main components, as a function of the output parameters of the
transmission, i.e. the output torque and speed of the hydraulic motor. The losses in the hoses and
pipelines between the hydraulic pump and motor are included in the overall efficiency results.

a) b)

.
0,0
o .
0,9

0.8 0,8

0,7 1
.0,7
't2
0,6 50.6 0
.1 Hyd motor
0,5 61 Nyd motor 0,5
4 2 El motor
.2. El. motor
93 Hyd pump
0,4

0,3
1:4
3THoYtdaI Pd7v e
P 7i 0,4

0,3
64 Total drive .

5 10 15 20 25 30 0 16000 20000 30000 40000 50000


Motor speed [rpm] Motor torque [Nm]

Fig. 2. Measured values and curve fits for the total efficiency of the transmission and its
components versus the motor speed and motor torque [3].
a). Measured efficiency at a motor torque of 48 kNm.
b). Measured efficiency at a motor speed of 25 rpm (pump dispi. setting approx. 0.86).

As can be seen, the efficiency depends on the output power level. In the investigated operating
parameter range the efficiency increases with an increased output power level from the transmission
and its components. However, a further increase in the output power level must not necessarily
imply an increase in efficiency. Wilson [4] has studied the performance of displacement pumps and
motors by the use of a dimensionless parameter. This parameter is defmed as the product of fluid

4
viscosity (µ) and speed (n) divided by the differential pressure (Ap), i.e. (u.n)/Ap. It is demonstrated
that the optimum total efficiency of a displacement machine is depending on the value of this
parameter. For instance, an increase in working pressure while holding viscosity and speed constant
will eventually result in a decrease in efficiency.

The estimated volumetric and hydromechanical efficiency of the hydraulic motor versus the motor
speed and motor torque is shown in Fig. 3.

a) b) [MPa
10 15 20 25
mu ____n

ei
0,98 0,98
alliMpilifflialrrr:

5.096 5. 0,96
.5
MRI
F P.
,94 50 0,94
0 1. Hydromech. eft
.2 Volumetric eft IM ] 01 ro h.elf.
0,92 0.92 2] Volumetric
c off]
111111111ffl ' ]
0,9 0,9
0 5 10 15 20 25 30 0 10000 20000 30000 40000 50000
Motor speed [rpm]
Motor torque [Ner!

Fig. 3. Estimated values and curve fits for the hydraulic motor's volumetric and hydromechanical
efficiency versus the motor speed and motor torque [3].
a). Hydromechanical and volumetric efficiency for the hydraulic motor at 48 kNm.
b). Hydromechanical and volumetric efficiency for the hydraulic motor at 25 rpm.

When the motor speed increases, the volumetric efficiency increases and the hydromechanical
efficiency decreases. With increased output torque, i.e. increased differential pressure, the
volumetric efficiency decreases and the hydromechanical efficiency increases. This is in accordance
with the results presented by Wilson [4], i.e. an increase in the dimensionless parameter implies an
increase in the volumetric efficiency and a decrease in the hydromechanical efficiency.

The estimated volumetric and hydromechanical efficiency of the hydraulic pump is shown in Fig. 4.

a) b)
Pump dispi. setting np IMPal
0.2 0,4 0,6 0.8 5 10 15 25

0.95 0,95

0.9 0,9
5.
0,85 e!, 0,85
tt LT,
0.8 0,8
Hydromech. eff.
0,75 02 Volumetric eff. 0.75

0,7 0,7
5 10 15 20 25 30 10000 20000 30000 40000 50000
Motor speed Nan] Motor torque [him]

Fig. 4. Estimated values and curve fits for the hydraulic pump's volumetric and hydromechanical
efficiency versus the motor speed and motor torque [3].
a). Hydromechanical and volumetric efficiency for the hydraulic pump at 48 kNm motor
torque Op approx. 24.5 MPa).
b).Hydromechanical and volumetric efficiency for the hydraulic pump at 25 rpm motor
speed (pump stroke approx. 0.86).
5
When the differential pressure is increased, the volumetric efficiency decreases and the
hydromechanical efficiency increases, i.e. the same behaviour as for the hydraulic motor. With an
increased displacement setting at the constant rotational speed, both the volumetric and the
hydromechanical efficiency increase. Thus, the best total efficiency is obtained at displacement
setting 1 for the studied operating range.

2.2.2 Efficiency models

Models are used in studying the static and dynamic behaviour of physical systems and are of central
importance, e.g. in control engineering and in optimisation techniques. When optimising a hydraulic
machine or a hydraulic system with regard to efficiency, an accurate model of the physical
behaviour governing flow and torque losses is needed. According to [5] there are two approaches in
deriving a specific model. The first approach is to base the models on prior generalized experience
expressed by physical laws. The second approach is to develop empirical models based on the use
of experimental identification. Several different steady state loss models for hydraulic pumps and
motors have been developed and can be found in the literature. Some of these models have been
created using the physical nature of the losses, while others are more or less empirical.

Wilson [6] was one of the first to develop a model to explain the physical behaviour governing flow
and torque in hydraulic displacement pumps and motors. The Wilson model takes into consideration
the laminar leakage and constant leakage flows, as well as torque losses due to Coulomb, viscous
and constant friction. Schlösser [7] and Thoma [8] extended this linear model. Schlösser takes the
density-dependent flow and torque losses into consideration, and Thoma's torque model is for
variable displacement units.

However, investigations of the loss behaviour of different displacement units have revealed that
these linear theories with constant coefficients do not fully explain the appearance of the flow and
torque characteristics. Hibi and Ichikawa [9] have proposed a modification of the Wilson torque
model to describe the torque performance of hydraulic motors from start to maximum speed. Their
model involves a friction component that decreases with increasing speed. On the basis of the
Wilson model, Dorey [10] has evaluated a flexible modelling technique where coefficients for the
flow and torque losses are used and the variation of the losses with speed, pressure and
displacement is built into the description of these coefficients. The development of hydraulic pump
and motor models has also led to models, [2], [11], where each loss source is described with several
functions which in some cases do not have any direct physical significance.

There have also been fully empirical models proposed, [12], [13], whose development has been
based on experimental data. This type of model is very valuable when modelling steady state
behaviour for control engineering applications. In [13] empirical models of the individual
components in a mobile hydrostatic transmission are used to optimise the efficiency of the
transmission. The speed of the diesel engine and the displacement setting of the hydraulic pump and
motor are varied to obtain the most efficient drive mode at specific operating points. However, the
complexity of empirical models and their somewhat arbitrary form make it difficult to relate
particular loss coefficients to physical effects and difficult to make comparisons between models
representing different designs of pumps and motors [14].

6
2.2.3 Simulation tools

The coefficient models derived from the physical behaviour of the losses provide a helpful tool for a
qualitative understanding of the phenomena governing losses in hydraulic machines. However, for
optimisation of an individual component, it is necessary to go one step further and investigate the
individual loss sources in the component.

Sealing and bearing gaps are one of the essential design elements of displacement machines. The
gaps usually undertake two different functions: the function of a sliding bearing between parts
moving relatively to each other and a sealing function [15]. The losses in hydrostatic machines are
to a large extent influenced by the friction and leakage flow in the different lubricating gaps within
the machine [16]. Consequently, an optimisation of the machine means essentially optimisation of
these lubricating gaps [17]. The design of sealing and bearing gaps in hydrostatic machines is
normally based on experience of former products. For the computation of gap flow there have been
different simulation models presented, e.g. [18], [19], [20] and [21].

1600

1200

0
800

400

0 5 10 15 20 25 30 35
Radial gap height [gm]

Fig. 5. Simulation results using CASPAR for power losses between the piston and cylinder bushing
of an axial piston swash plate pump as a function of the radial gap height.
1. high pressure: 20 MPa, displ. setting: I, rot. speed: 1500 rpm.
2. high pressure: 10 MPa, dispi. setting: 1, rot. speed: 1500 rpm.
3. high pressure: 10 MPa, dispi. setting: 1, rot. speed: 1000 rpm.
4. high pressure: 10 MPa, dispi. setting: 0.85, rot. speed: 1500 rpm.

CASPAR [22] is a simulation tool developed at the Technical University of Hamburg-Harburg in


Germany. This tool allows the calculation of the losses due to viscous friction and leakage in the
connected lubricating gaps in the rotating group of axial piston machines of the swash plate type.
On the basis of the given design parameters of the rotating group, the individual gap heights
between the piston and cylinder, between the slipper and swash plate and between the cylinder
block and valve plate and their dependency upon the angular position of the shaft can be calculated.
Due to the small gap heights, laminar flow is assumed and the Reynolds equation is used for
calculation of the gap flow. To calculate the gap flow in the gaps, CASPAR solves the motion
equation for all the moveable parts of the rotating group. The aim of the evaluation of the motion
equation is to modify the positions of the components for given external forces, so that the system
will reach an equilibrium state. The program is run for a number of shaft revolutions until the

7
positions of the components converge to a cyclical movement with the same period time as the shaft
speed.

In the lubricating gaps, power losses arise due to leakage and friction between parts in relative
motion to each other. In order to improve the performance of displacement machines, it is of
interest to find gap geometries that give low power losses. For example, in the contact between the
piston and the cylinder bushing of an axial piston swash plate pump, it is interesting to find the
optimum gap height that gives the minimum power loss. However, this optimum gap height
depends on the operating parameters of the machine, i.e. the speed, pressure and swash plate angle.
This is demonstrated in Fig. 5, which shows the results of a simulation of the power loss in the gap
between the piston and the cylinder bushing of an axial piston pump using the simulation tool
CASPAR. The simulations were made in order to find the gap height giving the minimum power
loss at different operating conditions. As can be seen, the simulation results indicate that the
optimum gap height depends on the prevailing operating conditions.

2.2.4 Experimental steady state measurements

Due to the complex behaviour of losses in displacement machines, the efficiency models and
simulation tools must be supplemented with steady-state measurements of the machine under study.
The international standard ISO 4409 defines the test procedures, conditions and required
measurement accuracy for the determination of the steady state performance of displacement
machines. The calculation of the volumetric and hydromechanical losses requires the determination
of the derived displacement volume. One experimental method for this is defined in the
international standard ISO 8426 and another experimental method was developed by [23]. In [16] it
is shown that these two methods are not quite exact. The method according to ISO gives different
values for the derived displacement volume for different pressures, and when using the method
suggested in [23], the derived displacement volumes obtained for different speeds differ from each
other. According to [24] the use of the standard methods for determination of the derived
displacement volume, which does not represent a constant value but varies with the operating
parameters, gives excessively high values for the internal volumetric losses. Therefore, a new
method for the estimation of volumetric and hydromechanical losses is presented in [24] which does
not need the determination of the derived displacement volume. With the help of results obtained
using the simulation tool CASPAR, it has been seen that the internal volumetric losses in axial
piston machines of the swash plate type usually are very small, i.e. they can be neglected with the
exception of internal volumetric losses caused by a cross port. This fact has been used to determine
the volumetric and hydromechanical losses from measured steady state performance characteristics.

2.2.5 Hydraulic fluids

The first hydraulic systems developed at the end of the 18th century used water as a hydraulic fluid.
The development of more sophisticated hydraulic components led to higher demands on the
lubricating properties of the fluid and at the beginning of the 20th century the use of mineral-based
oils began. Due to its good lubricating behaviour and its anti-corrosive properties, mineral oil
became the dominant hydraulic fluid. However, the use of mineral oils in hydraulic systems always
involves a fire hazard in the event of a leakage. Fire-resistant hydraulic fluids have therefore been
developed to reduce this hazard. Fire-resistance depends to a large extent on the amount of
incombustible components in the fluid, e.g. the amount of water. Moreover, environmental concerns
have led to the development of environmentally adapted hydraulic fluids. These fluids can typically
be vegetable-based fluids or fluids based on synthetic esters. A fluid has to fulfil certain

8
requirements regarding its effect on the environment in order to be classified as an environmentally
adapted hydraulic fluid.

The main function of a fluid in a hydraulic system is to transfer energy. Apart from this, there are a
number of other desirable properties for a hydraulic fluid. Typical examples are:

• Good lubricating properties • Protection against corrosion


• Suitable viscosity at the operating • Rapid air-release and separation of
temperature water
• Small viscosity variation with temperature • Compatibility with the system's
and time materials
• Long lifetime • Environmentally acceptable
• Chemical stability • Fire-resistance
• Good filtration capacity • Good heat transfer properties.

The term "viscosity" denotes the inner friction or resistance to flow of a fluid, and viscosity is
dependent upon the shear velocity, the temperature and the operating pressure. The viscosity of a
fluid normally increases with an increase in the shear stress and pressure, but decreases with an
increase in the temperature. A commonly used expression for the temperature-dependence of
hydraulic fluids is the viscosity index (VI). In a moving fluid the shear stress is often proportional
to the shear velocity. The fluid is then called a Newtonian fluid and the constant of proportionality
is named the dynamic viscosity. In hydraulics the term "kinematic viscosity" is often used. The
kinematic viscosity is obtained by dividing the dynamic viscosity by the density of the fluid.

1.0

•C 0.5

0
50 100 1000
Viscosity (cSt)

Fig. 6. Efficiency as a function of the kinematic viscosity. th, = volumetric efficiency,


rim = (hydro)mechanical efficiency, i, = total efficiency [25].

The general behaviour of the efficiency of a hydrostatic machine as a function of the kinematic
viscosity is shown in Fig. 6. As can be seen, an increase in the kinematic viscosity implies a
decrease in the mechanical (hydromechanical) efficiency and an increase in the volumetric
efficiency. The optimal viscosity region denotes the area where the highest total efficiency of the
machine is achieved. In this region the best balance between the hydromechanical losses and the
volumetric losses is obtained. The optimal viscosity region can vary depending on the prevailing
operating parameters, i.e. the pressure, speed etc. Attention must also be paid to the fact that

9
extreme viscosity values might affect the lifetime of a hydraulic component. The upper and lower
viscosity limits therefore define the boundaries for a proper operation of the machine.

In Section 2.2.1 some results from laboratory measurements of a hydrostatic transmission were
presented. Measurements were also carried out at different fluid temperature levels in order to
investigate how the viscosity of the hydraulic fluid affected the efficiency performance. In Fig. 7
the measured total efficiency of the complete transmission, the hydraulic motor and the hydraulic
pump is shown for two different operating points. The results are presented as a function of the
measured fluid temperature at the charge pump inlet.

a) b)

• OV,

0,9 09
.nre. 82 e
e 0.8 ,°,.; 0,8

01 Hyd. motor
0 02 Hyd. pump
.23 Total dove

0,6 0.6
30 40 50 60 70 80 30 40 50 60 70 80
Inlet temperature °C Inlet temperature °C

Fig. 7. Measured values and curve fits for the total efficiency of the transmission and its
components versus the temperature [3].
a). Measured efficiency at a motor torque of 48 kNm and a motor speed of 7 rpm.
b). Measured efficiency at a motor torque of 48 kNm and a motor speed of 25 rpm.

Within the studied temperature range, a decreased viscosity favours the total efficiency of the
hydraulic pump at both operating points. For the hydraulic motor a decreased viscosity is
favourable at the higher speed, whereas at the lower speed it implies a decrease in efficiency.
Therefore, it appears that the two components set different demands on the viscosity.

Selecting the correct fluid can be very important for the efficiency and reliability optimisation of a
hydraulic system [26]. Apart from the viscosity and its dependency upon pressure and temperature,
it can therefore be important to take the following fluid properties into account:

• Fluid density
• Coefficient of dynamic friction
• Coefficient of static friction
• Bulk modulus

In [26] the efficiency of an industrial hydrostatic transmission used in a belt conveyor drive is
studied. The influence of three different fluids on the overall efficiency of the drive is investigated
experimentally. The compared fluids are one mineral oil, one synthetic polyalphaolefin and one
vegetable fluid, all of the same viscosity grade. The results from the study indicate that the chosen
fluid type may affect the overall efficiency of the hydrostatic transmission.

10
3 LUBRICATION AND WEAR
Wear is a phenomenon of great technological significance. It is one of the three mechanisms by
which equipment fails, the others being fracture and corrosion [27]. Consequently, wear has a
decisive influence on the lifetime and the function of a component.

Tribology is the science of friction, wear and lubrication of moving or stationary parts. It is
concerned with contacts between surfaces in relative motion with or without the presence of
lubricants. The lubricants can be solid, liquid or gas. In hydraulics the hydraulic fluid, apart from
transferring energy, normally serves as a lubricant for the contacts moving relative to each other
inside hydraulic machines.

3.1 Lubrication regimes

Lubricants are used to influence the friction and/or wear between two surfaces in contact. In order
to provide a machine element with a satisfactory life, the physical and chemical interactions
between the lubricant and the lubricated surfaces must be understood. Depending on the grade of
separation between opposing surfaces, different lubrication regimes occur. As an aid in determining
the prevailing lubrication regime, the dimensionless film thickness parameter A can be used [28].

h
A= (1)
VR2q,a +R2q,b

where h. = minimum film thickness [m]


Ria,a = rms surface finish of surface a [m]
Rq,b = rms surface finish of surface b [m]

The film thickness parameter is used to define the prevailing lubrication regime. The range of A for
the different regimes is:

A<1 Boundary lubrication


1<A<3 Mixed film lubrication
3<A Full film lubrication

These three lubrication regimes can be seen in Fig. 8, where the friction coefficient (µ) is plotted
against the film thickness parameter (A).

In the full film lubrication regime, the surfaces are completely separated by the lubricant film and
no asperity contacts occur. The film thickness depends to a large extent on the viscosity of the
lubricant. Also the relative speed of the surfaces and the applied load is of importance. In boundary
lubrication the lubricant does not separate the contacting surfaces and there is considerable asperity
contact. The contact lubrication mechanism is governed by the physical and chemical properties of
thin surface films of molecular proportions, typically 1-10 rim in thickness. Additives in the
lubricants have therefore an important role as wear reducers and as friction modifiers. The
properties of the bulk lubricant are of minor importance in the boundary lubrication regime. Mixed
lubrication consists of a mixture of boundary and full film effects.

11
B) C)

-3 A

A) Boundary lubrication B) Mixed lubrication I C) Full film lubrication


Surface a
Rg.a

r- 011 Lubricant
Ru.b
Surface b

Fig. 8. Lubrication regimes.

3.2 Wear mechanisms

Wear involves the loss of material from a contact surface. The loss of material can be the results of
a variety of wear mechanisms. Examples of wear mechanisms are:

• Abrasive wear
• Adhesive wear
• Surface fatigue

Abrasive wear is the result of hard particles cutting the surface and removing material. It can occur
in a two-body process with two surfaces in sliding contact, or a three-body process where hard
particles get trapped between the sliding surfaces. Adhesive wear arises due to the shearing of the
small welds that occur at asperity contacts between two sliding surfaces. In time, the repeated
formation and breaking of such contacts lead to the detachment of loose particles. Surface fatigue is
caused by repeated mechanical stress and occurs e.g. in rolling contacts and in hydrodynamically
lubricated bearings subjected to alternating loads. Surface fatigue is also caused by cavitation, a
phenomenon that can occur in e.g. hydraulic pumps.

3.3 Wear tests for hydraulic fluids

The wear protection properties of a hydraulic fluid are important, as it normally serves as a
lubricant for the different contacts inside hydraulic machines. Often hydraulic fluids are specified to
meet the requirements of different major hydraulic pump manufacturers. An industry-standard vane
pump test for indicating the wear characteristics of petroleum and non-petroleum hydraulic fluids
employs a Vickers V104C vane pump. There are at least three national standards based on the use
of this pump: ASTM D-2882, DIN 51389 and BS 5096 (IP281/77) [29]. The Vickers V104C-test
involves the vane pump being run for a specified time in specified (fixed) operating conditions. The
total weight loss of the vanes and the ring becomes the quantitative value of the wear. However, a
drawback with the standardised vane pump test (ASTM D-2882) is that it does not provide the
necessary correlation required for prediction of the lubricating properties in other pump

12
configurations [30, 31]. Apart from Vickers V104 C, there are other pump tests to assess the anti-
wear performance of hydraulic fluids, e.g. Denison HF-0 and Vickers 35VQ25 [29].

The FZG test according to DIN 51354 is also often used in the oil specifications for hydraulic
fluids. With the FZG gear tester or the FZG gear machine, the wear on a pair of spur gears is
measured at different load steps. When greater damage occurs, the test is stopped. The result is
stated in terms of the load stage reached when such damage occurs. According to [32] the
conditions or characteristics which it is possible to measure using this test are the extreme-pressure
properties, load carrying capacity, shear stability, wear etc. The FZG test provides, according to
[33], a forecast of wear that will occur in axial piston pumps. A similar correlation for vane pumps
and gear pumps has not been shown.

4 SUMMARY OF PAPERS
Paper A
The power losses in a full-size industrial hydrostatic transmission were studied experimentally in
[3]. In the transmission studied, the charge pressure is normally around 1.5-2.5 MPa. The study
demonstrated that the losses due to the charge pump contributed significantly to a low total
efficiency of the transmission at low and medium output power levels. Even without considering
the charge pump, the efficiency of the hydraulic pump was seen to lag behind that of the other main
components, i.e. the hydraulic motor and the electric motor, in the lower pressure range. Paper A
presents an investigation into the distribution of power losses in a variable axial piston swash plate
pump commonly used in stationary industrial hydrostatic transmission systems. The study focuses
on the low and medium pressure range, where it is known that many industrial hydraulic drive
applications are working. The churning losses in the pump were estimated experimentally by
measurements in a test rig. The leakage flow and the power losses in the contacts between the
piston and cylinder and between the slipper and swash plate respectively were simulated with the
help of the simulation tool CASPAR. For the pump studied, the churning losses were significant
under the operating conditions typically occurring in industrial drive applications. The simulation
results indicate that the leakage to the pump casing mainly originates from the gaps between the
pistons and their respective cylinders and between the slippers and swash plate. It can also be
concluded that a considerable part of the power losses in the studied pump can be attributed to other
loss sources than those studied in this paper.

Paper B
The experimental measurements presented in [3] showed that the hydromechanical losses in the
radial piston hydraulic motor, i.e. the torque losses, dominated in a large part of the studied
operating parameter range. The aim of Paper B was to study two sliding contacts inside the
hydraulic motor and investigate their influence on the torque and power losses. Moreover it was
investigated whether and when a change in the lubrication regime could be expected in these
contacts. This was accomplished by a combination of experimental and theoretical studies, with a
special focus on two lubricated sliding contacts: the distributor valve contact and the piston/cam
roller contact. The theoretical study was conducted by simulations using a commercially available
FEM software package. This software makes it possible to consider fluid film behaviour between
loaded sliding surfaces taking fluid properties into account. The distributor valve contact was
modelled as a hydrostatic annular multi-recess plane bearing which had different recess pressures
and was working against a plane surface. The prevailing film thickness was assumed to be equal to
the minimum film thickness where fluid film lubrication could be assumed. The piston/cam roller
contact was modelled and simulated as a journal sliding bearing. On the basis of these assumptions,

13
the theoretical analysis of the contacts indicates that the distributor valve contact to a relatively
large extent can contribute to the external leakage of the motor. The results also show that the
piston/cam roller contact can enter the mixed lubrication regime at low motor speeds.

Paper C
Paper C describes a comparative investigation into the friction properties and wear protection
abilities of different hydraulic fluids. At low running speeds, an increased wear rate has been noted
in the contact between the cylinder bore and the piston skirt in a radial piston hydraulic motor. It is
of interest to investigate how this wear is affected by the fluid type used, partly to see if there are
commercially available fluids which could improve wear protection at low motor speeds and partly
to try to predict which fluids can cause problems in future low speed drive applications. The aim of
this paper was therefore to investigate the possibilities of simulating this wear in a model test and to
study the influence of different types of commercially available hydraulic fluids on wear protection
and friction. To simulate the contact between the cylinder bore and the piston skirt in the hydraulic
motor, tests were performed in a reciprocating test rig where the contact geometry was of the
cylinder-plate type. In the model test a synthetic ester aimed at meeting the conditions in water
turbine applications received a top ranking regarding both friction and wear protection properties. It
was also concluded that the use of water glycols could cause increased wear in hydraulic motors
operating in low speed applications.

5 CONCLUDING REMARKS
A prerequisite for the good performance of a hydrostatic transmission system is that its main
components should work well together. Being able to predict the loss behaviour of the individual
components increases the possibilities of designing and operating them so that they will work
together in an energy-efficient way. Loss models and especially simulation tools constitute effective
means of accomplishing this. Paper A and B present initial investigations into this area.

It is also important to have a detailed knowledge about the demands set by individual drive
applications. These demands should be translated into requirements that concern not only the main
components but also the auxiliary equipment of the hydrostatic transmission. For instance, it would
be interesting to investigate the possibility of tailoring the charge pressure setting (and the charge
flow) depending on the prevailing operating parameters and the demands set by the individual drive
application.

In Paper C a synthetic ester shows good properties regarding both friction properties and wear
protection in a model test. An increased knowledge about the behaviour in the different lubricated
contacts inside hydraulic components makes it possible to set demands on individual fluid
parameters. Clearly, the development of new hydraulic fluids is likely to open up interesting
opportunities for energy-saving and performance-enhancing activities in hydraulic drive systems.

14
6 REFERENCES
[1] Ivantysyn, J., Ivantysynova, M., Hydrostatic Pumps and Motors: Principles, Design,
Performance, Modelling, Analysis, Control and Testing, Akademia Books International, New
Delhi, India, 2001, ISBN 8185522162.

[2] Rydberg, K-E., On performance optimization and digital control of hydrostatic drives for
vehicle applications, Linköping Studies in Science and Technology. Dissertation No. 99,
Linköping, Sweden, 1983, ISBN 9173726699.

[3] Olsson, H., Isaksson, 0., Experimental Investigation of Efficiency in an Industrial Hydrostatic
Transmission, Hydraulikdagar i Linköping, 18-19 maj 1999, Linköping, Sweden, 1999,
ISBN 9172195088.

[4] Wilson, W.E., Positive-Displacement Pumps and Fluid Motors, Pitman Publishing
Corporation, New York, USA, 1950.

[5] Brokhattingen, ET., Conrad, F., Soerensen, P.H., Trostmann E., Experimental Determination
of Static Characteristics of Hydraulic Motors and Pumps, 11th IFAC World Congress 1990,
Tallin, Estonia, Vol. 8, pp. 272-280, 1990.

[6] Wilson, W.E., Performance Criteria for Positive-Displacement Pumps and Fluid Motors,
ASME Semi-annual Meeting, Paper No. 48— SA-14, 1948.

[7] Schlösser, W.M.J., Mathematical Model for Displacement Pumps and Motors, Hydraulic
Power Transmission, pp. 252-257 and 269, April 1961, pp. 324-328, May 1961.

[8] Thoma, J., Mathematical Models and Effective Peiformance of Hydrostatic Machines and
Transmissions, Hydraulic Pneumatic Power, pp. 642-651, November 1969.

[9] Hibi, A., Ichikawa, T., Torque Performance of Hydraulic Motor in Whole Operating
Condition from Start to Maximum Speed and Its Mathematical Model, 4th International Fluid
th th
Power Symposium, pp. B3-29-B3-38, University of Sheffield, England, April 16 -18 1975.

[10] Dorey, RE., Modelling of Losses in Pumps and Motors, 1st Bath International Fluid Power
Workshop, pp. 71-97, University of Bath, UK, 1988.

[11] Zarotti, G.I., Nervegna, N., Pump Efficiencies Approximation and Modelling, 6th Int. Fluid
Power Symposium, Paper C4, pp. 145-164, Cambridge, UK, April 1981.

[12] Conrad, F., Trostmann, E., Zhang, M., Automatic Computer Controlled Bi-Directional
Dynamometer Applied for Identification of Static Performance and Experimental Modelling
of Losses in Hydraulic Pumps and Motors, 3rd Scandinavian International Conference on
Fluid Power, pp. 129-148, Linköping, Sweden, 25-26 May 1993, ISBN 9178710898.

[13] Huhtala, K., Modelling of Hydrostatic Transmission — Steady-State, Linear and Non-Linear
Models, Acta Polytechnica Scandinavica, Mechanical Engineering Series No. 123, Helsinki,
Finland, 1996, ISBN 9526664997.

[14] McCandlish, D., Dorey, R.E., The Mathematical Modelling of Hydrostatic Pumps and
Motors, Proc. Instn. Mech. Engrs, No. 10, Vol. 198B, pp. 165-174, 1984.

15
[15] Ivantysynova, M., A New Approach to the Design of Sealing and Bearing Gaps of
Displacement Machines, 4th JHPS International Symposium, pp. 45-50, Tokyo, Japan, 1999,
ISBN 4931070043.

[16] Ivantysynova, M., Ways for Efficiency Improvements of Modern Displacement Machines, 6ffi
Scandinavian International Conference on Fluid Power, SICFP'99, pp. 77-92, Tampere,
Finland, 1999, ISBN 9521501812.

[17] Lasaar, R., The Influence of the Microscopic and Macroscopic Gap Geometry on the Energy
Dissipation in the Lubricating Gaps of Displacement Machines, 1st FPNI-PhD Symposium,
pp. 101-116, Hamburg, Germany, 2000, ISBN 3000065105.

[18] Fang, Y., Shirakashi, M., Mixed Lubrication Characteristics between the Piston and Cylinder
in Hydraulic Piston Pump-Motor, Transactions of the ASME, Vol. 117, pp. 80-85, January
1995.

[19] Kleist, A., Design of Hydrostatic Bearing and Sealing Gaps in Hydraulic Machines, 5th
Scandinavian International Conference on Fluid Power, pp. 157-169, Linköping, Sweden,
1997, ISBN 9178719755.

[20] Zhu, D., Cheng, H.S., Arai, T., Hamai, K., A Numerical Analysis for Piston Skirt in Mixed
Lubrication — Part land II, Journal of Tribology, Vol. 114, pp 553-562, July 1992, Vol. 115,
pp. 125-133, January 1993.

[21] Wieczorek, U., Ivantysynova, M., Computer Aided Optimization of Bearing and Sealing Gaps
in Hydrostatic Machines — a CAE-tool for Design Engineers, Proceedings 1st Bratislavan
Fluid Power Symposium, Bratislava, Slovakia, 1998.

[22] Wieczorek, U., Ivantysynova, M., Computer Aided Optimisation of Bearing and Sealing Gaps
in Hydrostatic Machines — the Simulation Tool CASPAR, International Journal of Fluid
Power, No. 1, Vol. 3, pp. 7-20, 2002.

[23] Toet, G., Die Bestimmung des theoretischen Hubvolumens von hydrostatischen
Verdrängerpumpen und -motoren aus volumetrischen Messungen, Ölhydraulik und
Pneumatik 14, Nr. 5, pp. 185-190, 1970.

[24] Ivantysynova, M., Energy Losses of Modern Displacement Machines — a New Approach of
Modelling, 7t Scandinavian International Conference on Fluid Power, SICFP 01, Linköping,
Sweden, 2001.

[25] Hodges, P., Hydraulic Fluids, Arnold, London, UK, 1996, ISBN 0340676523.

[26] Dahldn, L., Efficiency of Hydraulic Systems Influenced by Fluid Properties, Department of
Mechanical Engineering, Division of Machine Elements, 1998:38, Luleå, Sweden, 1998,
ISSN 14021757, ISRN LTU-LIC-98/38.

[27] Summers-Smith, J.D., An Introductory Guide to Industrial Tribology, Mech. Eng.


Publications Ltd, London, England, 1994, ISBN 0852988966.

[28] Hamrock B. J., Fundamentals of Fluid Film Lubrication, International Editions, McGraw-Hill
Inc., USA, 1994, ISBN 0070259569.

16
[29] Bishop, R. J. Jr., Totten, G.E., Tribological Testing with Hydraulic Pumps: a Review and
Critique, Tribology of Hydraulic Pump Testing, ASTM STP 1310, 1996, ISSN 00660558.

[30] Melief, H. M., Proposed Hydraulic Pump Testing for Hydraulic Fluid Qualification,
Tribology of Hydraulic Pump Testing, ASTM STP 1310, 1996, ISSN 00660558.

[31] Young, K. J., Hydraulic Fluid Wear Testing and Development, Tribology of Hydraulic Pump
Testing, ASTM STP 1310, 1996, ISSN 00660558.

[32] Mäkelä, K. K., Methods and Apparatuses for Tribological and Rheological Testing,
TRIBOLOGIA — Finnish Journal of Tribology, No. 2, Vol. 17,1998.

[33] Reichel, J., Pump Testing Strategies an Associated Tribological Considerations — Vane Pump
Testing Methods ASTM D 2882, IP281 and DIN 51389, Tribology of Hydraulic Pump
Testing, ASTM STP 1310, 1996, ISSN 00660558.

17
Paper A
POWER LOSSES IN AN AXIAL PISTON PUMP USED IN
INDUSTRIAL HYDROSTATIC TRANSMISSIONS

Håkan Olsson
ITH, The Institute for Applied Hydraulics
Sörbyvägen 1
891 61 Örnsköldsvik, Sweden
Phone +46 660 79864, Fax +46 660 79859
E-mail: hakan.olsson@ith.se

ABSTRACT

This paper presents an investigation into the distribution of power losses in a variable
axial piston swash plate pump commonly used in stationary industrial hydrostatic
transmission systems.

The churning losses in the pump have been estimated experimentally by measurements
in a test rig. The leakage flow and the power losses in the contacts between piston and
cylinder and between slipper and swash plate respectively were simulated with the help
of the simulation tool CASPAR.

For the pump studied, the churning losses are significant under the operating conditions
typically occurring in industrial drive applications. The simulation results also indicate
that the leakage to the pump casing mainly originate from the gaps between the pistons
and their respective cylinders and between the slippers and swash plate.

KEYWORDS: Axial piston pump, Hydrostatic transmission, Power loss, Simulation

NOMENCLATURE

x, y, z cartesian co-ordinates [m]


vx fluid velocity in x-direction [m/s]
vY fluid velocity in y-direction [m/s]
h gap height [m]
hG gap height between slipper and swash plate [m]
time [s]
Ff fluid force [N]
F, external force [1'1]
F, contact force [N]

Al
E Youngs modulus [Pa]
a normal stress [Pa]
z relative deformation [-]
11 dynamic fluid viscosity [Pas]
P pressure [Pa]
Ph high pressure [Pa]
Pi low pressure (charge pressure) [Pa]
Ap differential pressure (ph-pt) [Pa]
Vin kinematic fluid viscosity at the pump inlet (low pressure side) [rn2/s]
a pump displacement setting [-]
w rotational speed [rad/s]
Qext measured external leakage from the pump casing [1113/s]
Qext/Ap measured external leakage divided by diff. pressure [m3/Pas]
Qp simulated leakage between piston and cylinder [m3/s]
Qs simulated leakage between slipper and swash plate [m3/s]
Qp+Qs sum of simulated leakage between piston and cylinder and [m3/s]
between slipper and swash plate
Tin measured input torque to the hydraulic pump [Nm]
Tin/Ap measured input torque divided by diff. pressure [Nm/Pa]
Pout measured hydraulic output power of the hydraulic pump [W]
Pitot total measured power loss of the hydraulic pump [W]
P, measured power loss due to churning in the fluid filled pump casing [W]
Pp simulated power loss between piston and cylinder [W]
Ps simulated power loss between slipper and swash plate [W]
Pp+13, sum of simulated power loss between piston and cylinder and [W]
between slipper and swash plate
°casing fluid temperature in the casing [K]
Ogapmax maximum fluid temperature in the gaps [K]

1 INTRODUCTION

This paper presents an investigation into the distribution of power losses in a variable
axial piston swash plate pump commonly used in stationary industrial hydrostatic
transmission systems. Industrial low speed- high torque hydrostatic drives are
commonly used in continuous operation in a wide range of drive applications, e.g.
shredders, crushers, conveyors and cement kilns. The demands placed on such drives
change from application to application. In many cases the system is designed so that the
high-pressure level during normal operation is held relatively low (10-15 MPa) due to
the fact that a large starting (or emergency) torque is required. Additional requirements,
such as extended working life or high availability, make it sometimes necessary to
dimension the drive to work at pressure levels below that which the system may
ultimately be capable of. The hydrostatic drives used in cement kilns, for example, often
work continuously in their lower pressure range. In many situations, hydrostatic
transmissions have to compete with electromechanical drives, such as DC or AC
electric motors combined with gearboxes. In such situations, energy efficiency is a key
selection criteria where even a small increase in the efficiency of high power industrial
drives would give substantial savings.

A2
The power losses in a full size industrial hydrostatic transmission have been studied
experimentally by the author [1]. In the transmission studied, the charge pressure is
normally around 1,5-2,5 MPa. The study demonstrated that the losses due to the charge
pump contributed significantly to a low total efficiency of the transmission at low and
medium output power levels. Even without regard to the charge pump the efficiency of
the hydraulic pump was seen to lag the other main components, i.e. the hydraulic motor
and the electric motor, in the lower pressure range. It is therefore of interest to further
investigate the performance of this unit with a particular emphasis on the low and
medium pressure range where it is known that many industrial hydraulic drive
applications are working.

The term efficiency is an important parameter in power transmitting machines.


Efficiency describes the relationship between the ideal, loss-less working process and
the working process including losses. Several papers have dealt with the loss behaviour
of displacement machines. Wilson [2] was one of the first to try to explain the physical
behaviour governing flow and torque in displacement machines, taking into
consideration the laminar leakage flow in clearances as well as torque losses due to
coulomb, viscous and constant friction. Schlösser [3] extended this linear theory by
taking density dependent flow- and torque losses into consideration. Thoma [4]
presented a torque model useful for variable displacement units which included the
displacement setting when describing the hydrodynamic loss torque. Investigations into
the loss behaviour of different displacement units have, however, revealed that linear
theories with constant coefficients do not fully explain the flow- and torque
characteristics. These linear models have therefore been extended by a number of
authors to include various additional terms and to accommodate non-linearities [5], [6],
[7], [8] and [9]. However, these models do not relate particular loss coefficients to
physical effects and their use for predicting possible improvements in the loss behaviour
of a specific component is therefore rather limited.

Sealing and bearing gaps are essential design elements in displacement machines. In
Fig. 1 the main gaps in an axial piston pump are shown. Gaps usually undertake two
different functions: acting as a sliding bearing between parts moving relatively to each
other and to accomplish a sealing function [10]. The losses in hydrostatic machines are
to a large extent influenced by the friction and leakage flow in the different lubricating
gaps within the machine [11]. The design of sealing and bearing gaps in hydrostatic
machines is normally based upon experience gained from existing products [12]. A
number of different simulation models to help calculate gap flow have been presented
[13], [14], [15] and [16]. CASPAR [17] is a simulation tool developed at the Technical
University of Hamburg-Harburg in Germany. This tool calculates the individual gap
heights in the connected lubricating gaps of the rotating group in swash plate type axial
piston machines. Based on given design parameters of the rotating group the losses due
to viscous friction and leakage in the gaps between piston and cylinder, between slipper
and swash plate and between cylinder block and valve plate respectively can be
calculated. The simulation tool has been validated by comparing the sum of the
simulated gap leakages with experimentally measured leakage flow from the pump
casing [12]. From a scientific point of view, it is interesting to investigate whether this
tool can be used for a swash plate axial piston pump of a larger size and a different
make and study if the simulated distribution of leakage between the gaps shows a
similar behaviour. Results presented in [17] indicated that the simulated leakage
between cylinder block and valve plate contribute only around 3% of the total simulated

A3
leakage in the studied operating point. It would therefore appear that most of the
simulated leakage results from the gap flows between the piston and cylinder and
between the slipper and swash plate.
Churning losses result
cylinder block i piston ball bearing slipper
from the motion of
valve plate cylinder piston swash plate the rotating parts of
the displacement
machine in the fluid
filled casing. The size
of this torque loss
increases with
increasing rotational
speed of the machine.
According to [18]
these losses cause
noticeable power loss
in axial piston pumps
when idling, but are
not of much
Fig. 1 Sealing and bearing gaps in an axial piston pump [16]
significance under
normal operating conditions. A simplified theoretical reflection related to the churning
losses in an axial piston pump is presented in [19] where it is assumed inter aha that the
torque loss due to churning is proportional to the cube of the pitch radius of the cylinder
block. Thus, in larger pumps running at relatively high speed where the continuous
working pressure is low, the magnitude of the churning losses might be significant.

In this study the churning losses in the pump have been estimated experimentally by
measurements in a test stand. The leakage flow and the power losses in the contacts
between piston and cylinder and between slipper and swash plate respectively were
simulated with the help of the simulation tool CASPAR. The contribution of these
losses to the total measured power losses of the pump was then estimated in certain
operating points. The study focussed on the low and medium working pressure range
with the pump speed is fixed at approx. 157 rad/s.

2 METHOD AND MATERIALS

2.1 Experimental set-up

A schematic drawing of the experimental set-up and the position of the sensors are
shown in Fig. 2. The sensors and their location conform to ISO 4409, accuracy class B.
The fluid used was a mineral oil, ISO VG 100.

Two hydraulic pumps were mounted on the same shaft to form a regenerative test
system. The theoretical displacement volume of both pumps was 39,8 cm3/rad. The
pump under test (1) was a swash plate variable axial piston pump designed for closed
loop operation with a rated speed of 188,5 rad/s and a rated pressure of 35 MPa. This is
the same pump that was studied in [1]. In normal use, the tested pump would be

A4
equipped with charge- and pilot pumps mounted in tandem with the main pump. The
valves necessary for the closed loop operation, i.e. charge pressure relief valve, flushing
block and pressure override, are normally integrated in the main pump assembly.
However, the charge- and pilot pumps as well as the valve blocks necessary for the
closed loop system were removed from the test pump as only the main pump with its
rotating group were to be investigated.

Fig. 2 Schematic diagram of the test equipment.

An external boost pump (3) fed the closed circuit. This pump has a theoretical
displacement volume of 12,6 cm3/rad and runs at 157 rad/s. The pressure relief valve (4)
set the low-pressure level. A second pressure relief valve (5) was installed for safety
reason to limit the maximum high-pressure level. An adjustment screw controlled the
displacement setting of the test pump mechanically. The accuracy of the swash plate
angle at reduced displacement setting was ± 1%. The open loop pump (2) was equipped
with a pressure regulator and worked in motor mode to achieve a variable load pressure.
By adjusting the pressure regulator the high pressure level could be varied.

2.2 The simulation tool

To simulate the losses in the rotating group in the pump, CASPAR, a CAE design tool
for axial piston machines was used. As already mentioned, CASPAR is developed at the
Technical University of Hamburg-Harburg in Germany and allows the calculation of the
losses due to viscous friction and leakage in the connected lubricating gaps in the
rotating group of axial piston machines of swash plate type. Based on given design
parameters of the rotating group the individual gap heights between piston and cylinder,
between slipper and swash plate and between cylinder block and valve plate and their
dependency upon the angular position of the shaft can be calculated. Due to the small
gap heights, laminar flow is assumed and Reynolds equation is used to calculate the gap
flow. Eq. (1) shows the equation used for the piston/cylinder contact.

A5
a [ap 1-1 3 a ( ap 67 ah an 2 an)
+ (v ) +(vy) + (1)
at

For the contacts between slipper and swash plate and between valve plate and cylinder
block the Reynolds equation for a cylindrical co-ordinate system is used. To calculate
the flow in the gaps, CASPAR solves the motion equation for all moveable parts of the
rotating group. The motion equation in CASPAR is defined as:

— F, (h, t) — Fe — Fe (h, t) = 0 (2)

The fluid forces (Ff) are obtained from Reynolds equation by integrating the pressure
field. The external forces (Fe) consist of pressure forces, viscous friction forces, inertial
forces and centrifugal forces. If the lubricating film is interrupted, contact between the
sliding surfaces can occur. Therefore, a minimal gap height of 0,1 pm is assumed so
that Reynolds equation still can be applied for the calculation of the pressure field.
Where the gap height is smaller than the minimal allowed, elastic deformation of the
component is assumed. The normal stress (a) is calculated using the relative
deformation (c) and Youngs modulus (E) according to eq. 3.

a = EE (3)
The contact force (Fe) is calculated by integrating the normal stress. The aim of
evaluating the motion equation is to modify the positions of the components due to
given external forces, so that the system reaches an equilibrium state. The program is
run for a number of shaft revolutions until the positions of the components converge to
a cyclical movement with the same period time as the shaft speed.

The program requires the machine design parameters, i.e. the geometry of the different
gaps, as input. To obtain this data, the pump under study was disassembled and the
dimensions of the components in the rotating group measured. CASPAR also needs the
operating parameters as input, most of which can be obtained from steady state
measurements using sensors positioned as defined in ISO 4409. However, there are two
input parameters that are not easily measured. One is the relative rotation, which defines
the rotation of the slipper and piston relative to the cylinder block and the other is the
maximum temperature in the gaps. Olems [20] studied the temperature distribution in
the gap between piston and cylinder in an axial piston pump and results from this work
were used to estimate the maximum gap temperatures fed into CASPAR.

Output data from the program includes e.g. gap heights, gap flow, pressure fields and
viscous friction in the gaps, swash plate moment etc. The simulated external leakage
flow was compared with experimentally measured drainage flow rate to ensure that the
obtained simulation results were reasonable. In the experimental set-up all external
equipment is removed from the test pump. The measured leakage from the pump casing
therefore corresponds to the external leakage of the rotating group and can be used for
direct comparison with results from the simulation tool.

A more comprehensive description of the simulation tool CASPAR can be found in


[12], [16] and [17].

A6
3 RESULTS

In this section some results from the experimental measurements and the simulations
will be presented. The rotational speed of the pump during measurements was 156,7
rad/s (± 0,04%). In the simulations a value of 156,7 rad/s was used.

3.1 Result of experimental measurements

Fig. 3 shows the measured characteristics of the pump as a function of the high pressure
level under the following operating conditions:

Low-pressure level (pi): 2,5 MPa (± 2%)


Displacement setting (a): 1 (± 0,04 %)
Kin, fluid viscosity (vi.): 34 mm2/s (± 3%)

The numbers within brackets indicate the variation between individual measurements.

a) b)

0.9 ---T—•T
0 • • 0 .7 .17

6
j 0,7 2
75
0,6

0,4
3 6 9 12 15 12 15
High pressure [MN] Hi6gh pressure (M6
Pal
c) d)
600 100


500 o
80
6 ö° o.
k e•
E 400
ä o. 60 1
1•
p00 * . •11Z z,gx , mc.x z
• . 40 e
200 Q . •
I e 1 lln 1

,, j_2. Tin/Ake! 20
100
ce

0 0
3 12 15
Hi6gh pressure [MP
9a] High
6 pressure [MPal
9

Fig. 3 a) Measured total efficiency of the pump.


b) Measured hydraulic output power and measured power losses.
c) Measured external leakage and measured external leakage divided by zip.
ch Measured input torque and measured input torque divided by Ap.

The measured total efficiency of the pump as a function of the high-pressure level is
presented in Fig. 3a whilst Fig. 3b shows the measured hydraulic output power and the
power loss as a function of the high-pressure level. Fig. 3c and 3d show the measured
external leakage and the measured input torque as a function of the high-pressure level

A7
respectively. The external leakage shows a non-linear behaviour and appears to have a
minimum at a high-pressure level of about 6 MPa. The input torque increases more or
less linearly with magnitude of the high-pressure level. Fig. 3c and 3d also show the
relative external leakage and the relative input torque, i.e. the external leakage and input
torque related to the prevailing differential pressure level. Both the relative external
leakage and the relative input torque increase with decreasing pressure level and thus
contribute to the appearance of total efficiency curve shown in Fig 3a.

3.1.1 Estimation of churning losses


As can be seen in Fig. 3d, a higher relative input torque is required when the pressure
level decreases. This can largely be explained by losses that were more or less constant
during the measurements, for example, churning losses. To estimate the size of the
churning losses in the pump casing, measurements were conducted with the following
operating conditions:

Differential pressure (Ap): 3 MPa (± 1 %)


Pump stroke setting (a): 0,75, 0,85, 1,0 (± 0,04 %)
Fluid temp. pump casing: approx. 348 K (± 0,06 %)
Fluid temp. pump inlet: approx. 313 K (± 0,3 %)
Fluid temp. pump outlet: approx. 315 K(± 0,3 %)

At each pump displacement setting the power losses were first estimated with fluid
filled casing. The pump casing was then drained to a separate vessel via a ball valve
mounted in the bottom of the pump casing and the power losses with the empty casing
were estimated. As the differential pressure and the temperature in the loop are, in
principal, constant, the difference in loss torque between fluid filled and empty casing is
assumed to be attributed entirely to the effect of churning losses. By subtracting the
measured average values of the power losses with fluid in the casing from the measured
average power losses without fluid in the casing, the churning losses could be estimated
and are shown in Table 1.

Table 1. Estimation of the magnitude of churning losses in the pump casing.


Dispi. setting (a) Power loss diff.* - estimation of Pc [kW]
0,75 1,20
0,85 1,23
1,0 1,28
* Average values

3.1.2 Measurements at reduced low-pressure level


Performance measurements were conducted on the test pump at different low-pressure
levels while the differential pressure level was kept constant under the following
operating conditions:

Diff. pressure level (Ap): 8,5 MPa (± 0,8%)


Displacement setting (a): 1 (± 0,04 %)
Kin, fluid viscosity (vin): 35 mm2/s (± 4%)

Results from these experiments can be seen in Fig. 4.

A8
a) b)
0,9
40 8,5

0,88 • 8,1
.7_30 1
•9 •
120
3 0,64 I

0,82
di 10
I Oext1 _ 6,9
. 2. Pilot

0,8 6,5
0 0,5 I 1,5 2 2,5 0,5 1,5 2 2,5 3
Low pressure [MPal Low pressure [MPaj

Fig. 4 Measurements at different low-pressure levels and constant zip.


a) Measured total efficiency.
b) Measured external leakage and power loss.

3.2 Result of computer simulations

The simulation work was focused on the piston/cylinder and slipper/swash plate
contacts and their effect on leakage and friction. Some of the operating conditions that
were measured experimentally and presented in Fig. 3 were simulated with CASPAR.
Fig. 5a presents the results of the simulated leakage from the piston/cylinder contact and
the slipper/swash plate contact respectively. Also shown is the sum of these two. For
comparison purposes, the sum of the simulated leakages is also presented in Fig. 5b
together with the experimentally measured external leakage. Fig. 5c shows the
simulated power losses in the two gaps where it is assumed that viscous friction is
present in the contacts. In Fig 5d the sum of the power losses in the piston/cylinder- and
slipper/swash plate contacts is presented together with the measured total power losses
and the estimated power losses due to churning.

3.2.1 Simulated gap height between slipper and swash plate


As can be seen in Fig. 5b, both the measured external leakage and the sum of the
simulated leakage increase below a high-pressure level of about 6 MPa. According to
the simulations, this behaviour can be attributed to an increase in the gap heights
between the slipper and swash plate. In Fig. 6, the simulated gap heights are shown for
two of the operating points presented in Fig. 5. The high pressure phase is over a shaft
angle of ]0, Id and the low pressure phase over a shaft angle of DI, 27d. The position of
the slipper is described by the gap height between the slipper and swash plate at three
points, hGi-hG3, located at 2it/3 rad relative to each other on the outer diameter of the
slipper. The clearance between the slippers and swash plate is limited by the slipper
hold down device, which thus influences the maximum gap height during the low-
pressure phase.

A9
a) b)
50

40

1220
10

3 6 9 12 15 3 6 9 12 15
c) High pressure IMPS] d) High pressure [MPa]

1,5 10

1: 1
2
2 _9

1- 0,5

3 6 9 12 15 6 9 12 15
High pressure [MPa] High pressure [MPa]

Fig. 5 Simulation results as a function of the high-pressure level.


pi: 2,5 MPa, a: 1, v,n: 34 mm2/s.
a) Simulated leakage between piston/cyl. and between slipper/swash plate.
b) Sum of simulated leakage and measured external leakage.
c) Simulated power losses between piston/cyl. and between slipper/swash plate.
d) Measured total power loss, sum of the simulated power losses and power loss
due to churning.

a) b)
30

20

2 3 4 6 2 3 4 6
Shaft angle [red! Shaft angle [red]

Fig. 6 Simulated gap heights between slipper and swash plate for one shaft revolution.
pi: 2,5 MPa, a: 1, v,n: 34 mm2/s.
a) Simulated gap heights at 4,5 MPa high pressure.
b) Simulated gap heights at 11 MPa high pressure.

A10
3.2.2 Leakage variation with piston/cylinder radial clearance

In the piston/cylinder contact, the


20 maximum gap height is naturally
limited by the maximum radial
clearance between the piston and
the cylinder bore. Simulations
were made to investigate the
influence of this radial clearance
on the leakage. The results are
presented in Fig. 7 as a function
0 of the percentage change from
-75 -50 -25 0 25 50 75
Change of radial clearance from meas. average value NO the average value of the
measured radial clearance for the
Fig. 7 Simulated leakage between piston pump under study.
and cylinder. ph: 10 MPa, pi:
1,5 MPa, a : I, yin : 34 mm2/s.

3.2.3 Influence of maximum gap temperature


For the simulation results presented in Fig. 5-7, the maximum gap temperature was set
to 3K above the prevailing casing temperature. Table 2 presents results of simulations
using three different values of maximum gap temperature. All other operating
conditions are identical, i.e. the simulated operating point is the same.

Table 2.Simulation results with different values of maximum gap temperature.


h: 11 MPa, : 2,5 MPa, Jr yin: 34 mm2/s.
Ogapmax Qp [CM3/S1 Qs [CM3/S]

°Casing+ 1 K 8,17 23,17


OCasing+3K 8,33 23,83
Ocasing+10K 9,17 24,17

PlstorYcyll.er Piston/ender
a) 9% 9%

Swper/swasn plate Slipperlswash Oele


IS% 18%

raPistonIcylinder SPistoNcylinder
0SliPPeriswash plate .SkPPer/swash plate
Churning OChumino
55% .3 Others pOthers
Others
60%

To. power loss: 6,0 kW Total power loss, 8,33 kW

Fig. 8 Estimated distribution of losses in the hydraulic pump.


a). High pressure (ph) 6,5 MPa.
b). High pressure (ph) 12,5 MPa.

All
3.2.4 Estimation of loss distribution in the hydraulic pump
Based on the results shown in Fig. 5d, an estimation of the loss distribution in the
hydraulic pump is presented in Fig. 8 for the following two operating points:
• High pressure level (ph): 6,5 MPa and 12,5 MPa.
• Low pressure level (pi): 2,5 MPa.
• Displacement setting (a): 1.

The average value of the experimentally measured power loss at these operating points
were 6,0 and 8,33 kW respectively.

4 ANALYSIS AND DISCUSSION

The simulation results presented in this study are valid for the piston/cylinder and the
slipper/swash plate and their effect on the leakage and friction is studied. A complete
simulation of all three gaps in the rotating group, i.e. piston/cylinder, slipper/swash
plate and valve plate/cylinder block, implied a large increase in simulation time. Due to
time constraints, the valve plate/cylinder block contact was therefore not dealt with in
this study.

When estimating churning losses it is assumed that the conditions in the lubricating
gaps are the same both with and without fluid in the casing. This assumption is based on
the fact that the casing was emptied for a very short time and that the output power level
was held low in the tests. However, a fluid filled casing might be necessary for the
contacts in the rotating group to operate properly and the real churning losses under the
prevailing operating conditions could differ from those estimated in this study. The fluid
in the case acts as a medium for transporting heat and therefore the temperature
conditions in the gaps may change when the casing is drained, despite a low output
power level. This can in turn affect the friction- and leakage characteristics in the
lubricating gaps. The lubrication between slipper and swash plate may be inferred if the
fluid is drained. Hooke [21] made extensive investigations into the lubrication of the
slippers. In measurements conducted on overclamped slippers, it was noted that the gap
heights were significantly reduced if the slipper was not fully flooded by fluid on the
swash plate, in other words, that fluid fed from the slipper orifice alone was inadequate
to lubricate the slippers. A similar behaviour could possibly occur between valve plate
and cylinder block when fluid is not present. In [22] it is stated that the power losses
due to churning in gear systems depend upon the kinematic viscosity of the fluid to the
power of 0,6. According to the authors this dependence is also likely in hydraulic
pumps. Thus, to limit the influence of churning losses in the pump, the fluid viscosity in
the pump casing should be kept as low as possible.

According to Fig. 4 a reduction in the charge pressure level results in a small efficiency
increase in the main pump, at least down to a low-pressure level of 0,7 MPa. Below this
level, the power losses tend to increase. This is most likely due to an increase in friction
in some of the contacts as the measured external leakage decreases more or less
proportionally to the low-pressure level. However, there are causes that oppose a charge
pressure reduction. In some drive applications, for instance shredders, the pressure level
can be very oscillating due to irregular loads and thus a certain minimum low pressure
level is needed to avoid problems with cavitation due to sudden pressure changes. The

Al2
required charge pressure is also dependent on a phenomena related to fluid transients at
the start of the suction phase. Edge et al [23] have studied cylinder pressure transients
within an axial piston pump under high-speed conditions both theoretically and
experimentally. They found that a pressure overshoot occurs at the bottom dead centre
(BDC) and a pressure undershoot occurs at top dead centre (TDC). This study revealed
that fluid momentum effects during the start of the suction phase at TDC may cause the
cylinder pressure to decompress (undershoot) to a level below the pump inlet pressure.
The level of this pressure undershoot determines the minimum charge pressure required.
In their study it was seen that the pressure undershoot increased with an increase in
speed, delivery pressure and displacement setting. To limit the occurrence of cavitation
at the start of the suction phase in an axial piston pump Harris et al [24] proposed a
valve plate equipped with a pre expansion volume (PEV). The idea of this was to reduce
the magnitude of sudden outflow from the cylinder into the suction port. Simulation
results presented in [24] showed that this arrangement would help to reduce the
occurrence of cavitation over a relatively wide operating range. As the pressure
undershoot decreases with a decreased delivery pressure, the possibility exists of
lowering the charge pressure level in drive applications with steady, near constant
torque characteristics, especially in transmissions running at low working pressure
levels. Also, the fact that the pump normally runs at a fixed speed favours the possibility
to design the valve plate in such a way as to limit the size of the pressure undershoots.

The sum of the simulated leakages between piston and cylinder and between slipper and
swash plate corresponds fairly well with the measured external leakage as seen in Fig.
5b. The simulation results show a similar behaviour to the measured leakage, i.e. an
initial decrease followed by an increase as the high-pressure level is increased. There is,
however, some divergence between the simulated and measured leakage. One reason for
this could be the maximum gap temperature that was used in the simulation. It was
determined in [20] that the temperature in the gap between the piston and cylinder is
related to the magnitude of the high-pressure level. In Table 2, simulation results with
different values of the maximum gap temperature were presented and the leakage shown
to increase with increased gap temperature. It is quite possible that some of the
difference between measured and simulated results can be attributed to a difference in
maximum gap temperature used in the simulation and the real gap temperature. Another
reason for the difference between measured and simulated leakage could be slight errors
in measuring the pump geometry, e.g. the diameter of the pistons and cylinder bushings.
In CASPAR the geometrical dimensions of one piston is specified, which is then taken
as representing all the pistons in the rotating group. The average value of the measured
radial clearances of all nine piston/cylinder bushings was used, even though a variation
of up to 20 % from this average clearance value was noted between different pairs of
piston/cylinder bushing. As was seen in Fig. 7, the radial clearance can have a
significant influence on the leakage flow in this contact. Simulations were also made at
zero differential pressure. Due to time constraints, experimental validation at this
operating point has not yet been made. The design of the test rig limited the minimum
attainable differential pressure level.

Estimated power losses due to churning and power losses in the piston/cylinder contact
and the slipper/swash plate contact respectively are shown in Fig. 8. The figures for the
latter two contacts are based on the assumption that viscous friction is present in the
contacts. As can be seen, the sum of these three sources of losses constitutes less than
50 % of the total power losses over the operating parameter range studied. A large part

A13
of the power losses must therefore come from other loss sources, which have been
grouped under the name "others". The power losses in this group can be attributed to
the following:
• internal leakage and compression losses
• friction losses in bearings and seals
• pressure drop due to fluid flow through internal resistances such as in the
pressure ports
• friction and leakage in the valve plate/cylinder block contact.
The indications on low leakage from the valve plate/cylinder block contact imply small
gap heights. This in turn, can lead to high frictional losses in this contact. It can be seen
from the results presented in this paper that simulations of all three gaps in the rotating
group would give more information regarding the loss distribution in the pump. If the
complete rotating group is simulated, also the internal leakage and the compression
losses can be estimated by the use of CASPAR.

5 CONCLUSIONS

For the pump under study the following conclusions can be drawn:
• The churning losses are not negligible at the prevailing operating conditions.
• It would be interesting to investigate the possibility of tailoring the charge
pressure setting depending on the operating parameters and the demands set by
the drive application.
• CASPAR can be used to simulate the piston/cylinder- and the slipper/swash
plate contacts of the investigated pump in the operating parameter range under
study. As a next step it would be interesting to make simulations of the complete
rotating group to make predictions about internal leakage, compression losses
and losses between valve plate and cylinder block.
• The simulation results indicate that the experimentally measured external
leakage from the pump casing originates mainly from the gaps between piston
and cylinder and between slipper and swash plate.

REFERENCES

[1] Olsson, H., Isaksson, 0., Experimental Investigation of Efficiency in an Industrial


Hydrostatic Transmission, Hydraulikdagar, Linköping, Sweden, 1999,
ISBN 91-7219-508-8.
[2] Wilson, W.E., Performance Criteria for Positive-Displacement Pumps and Fluid
Motors, ASME Semi-annual Meeting, Paper No. 48-SA-14, 1948.
[3] Schlösser, W.M.J., Mathematical Model for Displacement pumps and Motors,
Hydraulic Power Transmission, April 1961, pp. 252-257 and p. 269, May 1961,
pp. 324-328.
[4] Thoma, J., Mathematical Models and Effective Performance of Hydrostatic
Machines and Transmissions, Hydraulic Pneumatic Power, Nov.1969, pp. 642-
651.
[5] Dorey, R.E., Modelling of Losses in Pumps and Motors, Bath International
Fluid Power Workshop, University of Bath, UK 1988.

A14
[6] Zarotti, G.I., Nervegna, N., Pump Efficiencies Approximation and Modelling, 6th
Int. Fluid Power Symposium, Cambridge, UK, 1981, paper C4, pp. 145-164.
[7] Rydberg, K-E., On Performance Optimization and Digital Control of Hydrostatic
Drives for Vehicle Applications, Linköping Studies in Science and Technology,
Dissertation No. 99, Linköping, Sweden, 1983.
[8] Conrad, F., Trostmann, E., Zhang, M., Automatic Computer Controlled Bi-
Directional Dynamometer Applied for Identification of Static Performance and
Experimental Modelling of Losses in Hydraulic Pumps and Motors, Proc. 3rd
Scandinavian Int. Conference on Fluid Power, Vol. 1, Linköping, Sweden, 1993.
[9] Huhtala, K., Vilenius, M., Comparison of Steady-State Models of Hydraulic
Pumps, 5th Scandinavian Int. Conf. on Fluid Power, Linköping, Sweden, 1997.
[10] Ivantysynova, M., A New Approach to the Design of Sealing and Bearing Gaps of
Displacement Machines, Fourth JHPS Int. Symposium, Tokyo, Japan, 1999.
[11] Ivantysynova, M., Ways for Efficiency Improvements of Modern Displacement
Machines, Sixth Scandinavian Int. Conf. on Fluid Power, Tampere, Finland, 1999.
[12] Wieczorek, U., Ivantysynova, M., CASPAR-a Computer Aided Design Tool for
Axial Piston Machines, Proceedings of the Bath Workshop on Power
Transmission and Motion Control PTMC, University of Bath, England, 2000.
[13] Fang, Y., Shirakashi, M., Mixed Lubrication Characteristics Between the Piston
and Cylinder in Hydraulic Piston Pump-Motor, Trans. of the ASME, Vol. 117,
Jan. 1995.
[14] Kleist, A., Design of Hydrostatic Bearing and Sealing Gaps in Hydraulic
Machines, 5 th Scandinavian Int. Conf. on Fluid Power, Linköping, Sweden, 1997.
[15] Zhu, D., Cheng, H.S., Arai, T., Hamai, K., A Numerical Analysis for Piston Skirt
in Mixed Lubrication-Part I and II, Journal of Tribology, 114, 115 pp. 553-562,
pp. 125-133, 1992.
[16] Wieczorek, U., Ivantysynova, M., Computer Aided Optimization of Bearing and
Sealing Gaps in Hydrostatic Machines, Proceedings ist Bratislavan Fluid Power
Symposium, Bratislava, Slovakia, 1998.
[17] Wieczorek, U., Ivantysynova, M., Computer Aided Optimisation of Bearing and
Sealing Gaps in Hydrostatic Machines-the Simulation Tool CA SPAR,
International Journal of Fluid Power 3 (2002) No. 1 pp. 7-20.
[18] Ivantysyn, J., Ivantysynova, M., Hydrostatic Pumps and Motors: Principles,
Design, Performance, Modelling, Analysis, Control and Testing, Akademia Books
International, New Delhi, 2001.
[19] Jang, D-S., Verlustanalyse an Axialkolbeneinheiten, Aachen Techn. Hochsch.,
Diss, 1997, ISBN 3-89653-252-9.
[20] Olems, L., Ein Beitrag zur Bestimmung des Temperaturverhaltens der Kolben-
Zylinder-Baugruppe von Axialkolbenmaschinen in Schrägscheibenbauweise. VDI
Fortschritt-Berichte. Reihe 1 Nr. 348. Düsseldorf, 2001 (in German).
[21] Hooke,C.J., The Design of Slippers for Axial Piston Pumps, The Second Bath
International Fluid Power Workshop, September 21st-22nd, Bath, England, 1989.
[22] Bronshteyn, L.A., Kremer, J.H., Energy Efficiency of Industrial Oils, Tribology
Transactions 42:(4), pp. 771-776, Oct 1999.
[23] Edge, K.A., Darling, J., Cylinder Pressure Transients in Oil Hydraulic Pumps
with Sliding Plate Valves, Proc. ImechE, Vol. 200, No Bl, pp.45-54, 1986.
[24] Harris, R.M., Edge, K.A., Tilley, D.G., Reduction of Piston Pump Cavitation by
means of a Pre-expansion Volume, Fifth Bath International Fluid Power
Workshop, England, Sept. 1992.

Al5
Paper B
ANALYSIS OF TWO SLIDING CONTACTS INSIDE A RADIAL
PISTON HYDRAULIC MOTOR

Leon DAHLEN* and Håkan OLSSON**

*Mid Sweden University, Depaitiiient of Engineering, Physics and Mathematics


S-831 25 Östersund, Sweden
(E-mail: Leon.Dahlen@mh.se)
**The Institute for Applied Hydraulics
Sörbyvägen 1, S-891 61 Örnsköldsvik, Sweden
(E-mail: hakan.olsson@ith.se)

ABSTRACT

This paper presents an investigation of the torque losses in a hydraulic motor of radial piston type
commonly used in industrial hydrostatic transmissions. When the speed of the hydraulic motor is
decreased an increase of the torque loss can be measured. In order to investigate the reason for this
increase a combination of experimental and theoretical studies are performed with special focus on
two sliding contacts-the distributor valve and the piston-cam roller contact. The theoretical analysis
of the contacts reveals that the distributor valve contribute more to torque and power losses
compared to the piston-cam roller contacts and largely to the external leakage of the motor. At low
speeds the piston-cam roller contact will enter the mixed lubrication regime.

KEY WORDS

Hydraulic motor, Radial piston, Torque loss, Simulation

INTRODUCTION experience expressed by physical laws. The


second approach is to develop empirical
Models are used in studying the static and models based on the use of experimental
dynamic behaviour of physical systems and identification.
are of central importance in e.g. control Several different steady state loss models for
engineering and in optimisation techniques. hydraulic pumps and motors have been
According to [1] there are two approaches in developed and can be found in the literature.
deriving a specific model. The first approach Some of these models are modelled using the
is to base the models on prior generalized physical nature of the losses [2], [3] and [4],

Bl
others could be described as semi-empirical reached in some of the contacts inside the
[5], [6], [7], and [8] while still others are more radial piston hydraulic motor studied in this
or less empirical, e.g. [9] and [10]. paper. The aim of this paper is to study two of
When optimising a hydraulic machine with the contacts inside the motor and investigate
regard to efficiency an accurate model of the their influence on the torque and power
physical behaviour governing flow and torque losses. Also it will be investigated if and
losses is needed. The coefficient models when a change in lubrication regime can be
derived from the physical behaviour of the expected.
losses provide a helpful tool for a qualitative
understanding of the phenomena governing
losses in hydraulic machines. However, for TORQUE LOSSES IN THE
optimisation purposes it is necessary to go HYDRAULIC MOTOR
one step further and investigate the individual
loss sources in the component. This is The radial piston motor of cam-lobe design
especially true under operating conditions has an even number of pistons that operates in
where mixed- or boundary lubrication is sequences of working pressure and charge
likely to occur. There have been models pressure in a reciprocating motion. The piston
developed to consider effects of mixed is operated via a cam roller in contact with the
lubrication, e.g. [11] and [12]. These models curved track cam ring, which is fixed to the
use an average Reynolds equation in terms of motor casing. The casing is stationary, which
pressure and shear flow factors to include the means that the cylinder block and piston
effect of surface texture on partial assembly are rotating. In motor mode, the
hydrodynamic lubrication. piston assembly moves radially outwards
Under certain operating conditions it is under working pressure and inwards during
believed that a change in lubrication regime is idling. In Figure 1 a cross-section of the
motor is shown.

1. Cam ring
2. Cam roller
3. Cam roller bearing
4. Shaft bearing
5. Radial shaft seal
6. Guide plate
7. Piston
8. Valve plates
9. Shaft bearing
10.Radial shaft seal
11.Radial shaft seal
12.Shaft bearing

Figure 1 Cross-section of the hydraulic


motor.

B2
The torque losses are mainly caused by
different types of friction and the required
torque for impulse changes of the fluid
flowing through the motor.
Under steady-state conditions the hydro
mechanical losses are mainly believed to ,m>
occur in the following areas:
• Friction between the valve plates in
the distributor valve
• Friction between the piston ring and
piston bore and between the piston
outer surface and the piston bore
• Friction between piston and cam roller
• Rolling friction Figure 2. Schematic drawing of the test
,7 Between cam roller and cam ring equipment.
,7 In the roller bearings
•7 Between guide plates and cam In Figure 3 the torque losses as a function of
roller bearings differential pressure is shown for different
• Friction in radial shaft seals speeds.
• Pressure drop due to fluid flow 1000

through the motor


800
• Inertia losses of moving parts.
Torque loss/ Nm

BOO

In order to evaluate the total sum of torque


losses in the motor experimental tests were 400

performed. In Figure 2 a simplified schematic


drawing of the equipment is shown where
also the position of the sensors can be viewed.
The test circuit conforms to ISO 4409 10 15 20 25

accuracy class B. The test equipment consists Differential pressure/ MPa

of two separate hydraulic circuits where two Figure 3. Measured torque loss as a function
identical hydraulic motors are mounted on the of differential pressure at approx.
same shaft. The displacement of the motors is 100 cSt inlet fluid viscosity.
approx. 210 m3/rad. The hydraulic motor
under investigation (1) works in motor mode The torque losses increase with both speed
in a closed loop system where the low and pressure. The Reynolds numbers in the
pressure is set to approx. 2 MPa. The other operating conditions under study is low
motor (2) is working in pump mode towards a indicating that laminar flow would prevail in
pressure relief valve in order to accomplish the motor internal channels. However, the
the load torque. By adjusting the setting of the relationship between the length/diameter of
relief valve, the braking torque can be the channels and the corresponding Reynolds
changed. number indicates that fully developed laminar
The hydraulic fluid used was a synthetic flow will not occur, i.e. the length of the
polyalphaolefin (PAO) ISO VG 100. channels are much shorter than the transition
Measurements were conducted at different length. This implies that both viscous and
pressure-, speed- and fluid temperature levels. inertia forces will affect the pressure drop and
the friction factor will be larger than for pure

B3
laminar flow. Nevertheless, by regarding the model differ somewhat from the actual
flow as fully developed laminar flow geometry in order to simplify the model but
according to the Hagen-Poiseuille law and the area of the bearing load carrying surface
calculating with loss coefficients based on has remained unaltered. The dimensions of
data values from [13] the difference in torque the bearing are: inner radius (A) 77,5 mm,
loss between 7-13 rpm and 13-19 rpm outer radius (B) 93,5 mm, pad inner radius
respectively can to most extent be assigned to (C) 79,5 mm, pad outer radius (D) 91,5 mm
pressure drop due to fluid flow. and recess-pad angle (E) 14,7°. Both valve
At 1 and 3 rpm the torque losses are larger plates are assumed to be rigid in the model.
than at 7 rpm at a differential pressure of Theoretical analysis of bearing operations is
approx. 10 MPa, even though the pressure made by carrying out simulations on a model
drop due to fluid flows is substantially lower. created using a FEM software package
The torque losses at these two lower speeds developed in Sweden [14]. This package can
also show a larger pressure dependency and a analyse a fluid film contact taking fluid
smaller viscosity dependency than at higher parameters into account.
speeds. Referring to the Stribeck-curve one In Figure 5 the calculation flow chart is
can suspect that a change in lubrication shown using the supplied recess pressure as
regime is reached in some of the tribological input data. The outputs from the simulations
contacts inside the motor. Two tribological when convergence criteria are reached are
contacts where this might happen are between pressure and flow distribution, reaction
the valve plates in the distributor valve and in force/moment and power loss in the contact.
the contact between piston and cam roller. As The minimum value of film thickness for this
there seem to be few references dealing with bearing with regard to full film lubrication
the performance of these contacts they will be and surface roughness [15] is estimated to 15
studied more closely in the following. gm. The fluid flow temperature to the high-
pressurised pads was 40 °C and fluid flow
temperature to the low- pressurised pads was
SIMULATION OF DISTRIBUTOR 45 °C. The physical fluid properties were
VALVE analogous with the polyalphaolefin used in
the experimental tests.
The distributor valve provides the
distribution of oil to the piston bores. This
valve consists of two valve plates, one fixed
to the connection block and one moving with
the cylinder block. The valve plates are
pressed together by springs at start and by
balancing sleeves and balancing pistons
together with the springs under operating
conditions to control the separation between
the surfaces.
The contact between the two plates has been
simulated as a hydrostatic annular multi-
recess plane bearing with different recess
pressures. The bearing surface consists of Figure 4. Simplified model of the distributor
twelve recess-pads alternately pressurised by valve.
low and high pressure as shown in Figure 4.
The inside geometry of the recess-pads of the

B4
begins to develop. The diameter of the cam
Input of initial data: roller is 85 mm and the width of the cradle is
Pressure/load, temperatures, speed
and viscosity.
90 mm (see also Figure 1).

Calculate pressure distribution by


solving Reynolds equation.

4
Calculate velocity distribution in the
oil film

Calculate oil film temperature


distribution by solving Energy
equation

OLYWARC INWARD

Calculate
viscosity

Figure 6. Piston-cam roller assembly.

The same FEM package has been used to


analyse the hydrodynamic part of the contact
between the piston and the cam roller. The
contact have been modelled and simulated as
a journal slider bearing. The Reynolds
End
boundary conditions are used. The calculation
flow chart is shown in Figure 5 using the
Figure 5. Data flow chart during simulations. external load as input data. The outputs from
the simulations when convergence criteria are
reached are pressure and flow distribution,
SIMULATION OF PISTON-CAM film thickness, reaction force/moment and
ROLLER CONTACT power loss in the contact. Also here the
geometry of the model differ somewhat to the
Each piston (1) is, via a hydrostatic cradle, actual contact in order to simplify the model
pressed against a cam roller (2), which is but the area of the bearing load carrying
rotating on the cam ring (3), see Figure 6. The surface has remained unaltered. The
figure shows the piston movement during the minimum value of film thickness for this
stroke in driving mode (working pressure at bearing with regard to full film lubrication
outward movement). This contact is and surface roughness [15] is estimated to 5
hydrostatic balanced, but approximately 7 gm. The numbers of pistons in this motor are
percent of the pressure applied on the upper eight and under running conditions four of the
side of the piston has to be taken care of by pistons are supplied by high pressure and the
hydrodynamic bearing action. The hydrostatic other four pistons with low pressure due to
balance is accomplished by allowing retraining oil flow. The retraining pressure
pressurised fluid to pass through the piston to was 2 MPa in these simulations.
the sealing lands of the cradle. At start of the
motor this contact is pressed together and
separation of the surfaces occur when the cam
roller starts rotating and a hydrodynamic film

B5
RESULTS constant film thickness. The relationship in
leakage between the distributor valve
In Figure 7 the minimum film thickness of compared to piston-cam-rollers is
the piston-cam roller contact is shown at three approximately a factor of 5 times greater at 3
different motor speeds and at different high- rpm and 5 MPa and a factor of 200 times
pressure levels. The result shows that the film greater at 30 MPa at the same speed.
thickness decrease under the full film level of
25
5 gm at a speed of 1 rpm and a load of 10
MPa. 20

2,0E-05

Distnbutor at 7 ,m
; • Distributor at 3 rpm
1•Prstor Cam at 7 rpm
1,5E-05 ;
a • t3 Pis. Cars at 3 rpm

E • 7 rpm:
• 3 rpm
• 1 rpm
1,0.5
10 15 20
• Pnewure/ mva

5.0E -06

Figure 8. Torque losses at two different


0.0E. speeds.
15 20
Pressure/ MPa
500

Figure 7. Minimum film thickness in the 400

piston-cam roller contact at


different speed and load. 300

• Destnbutor at? rpm

k 200
o Distributor at 3 rpm

Figure 8 shows the torque losses from the • Piston Cam at 7 rprn
o Aston Cam at 3 rpm
contacts of eight piston-cam rollers under tOo

running conditions. Torque losses from


piston-cam rollers are very low compared to
15 20
torque losses of the distributor valve in the Pressure/ MPa

full film lubrication regime. Theoretical


Figure 9. Power loss at different speed and
power losses for the two contacts are shown
load.
in Figure 9. The contact of the distributor
valve creates more power losses compared to
the piston-cam roller contacts over the
DISCUSSION
presented simulated range. The differences of
power loss between these two contacts can be
The simulations show it is likely that a
related to the differences in leakage. The
change in lubrication regime occurs in the
leakage flow increase with pressure for the
piston-cam roller contact at rather low
distributor valve but decrease with pressure in
working pressures when the speed is
the piston-cam roller contacts. This is
decreased. However, as the software is
explained by the fact that for the piston-cam
limited to full film lubrication simulation it is
roller contact, which is simulated as a slider
journal bearing, the film thickness will difficult to estimate the size of the torque
losses in mixed and boundary lubrication
decrease when the load increase. In the
regimes. These losses could be estimated by
distributor case the contact is simulated as a
using an approach similar to [11] and [12].
hydrostatic multi-recess plane bearing with

B6
As long as full film lubrication is present the more significant influence on the friction and
torque losses from the two studied contacts is thereby on the losses.
small and thus do not explain the measured
increase in torque loss when the speed is
reduced from 7 to 3 rpm. This implies that CONCLUSIONS
other sliding contacts should be studied from
the aspect of mixed friction. A contact that • At low speeds the piston-cam roller
probably has a strong impact in this respect is contact will enter the mixed
the piston-cylinder contact. lubrication regime.
The variation and size of the calculated • The distributor valve contributes more
leakage flow from the two studied contacts to torque and power losses compared
with regard to the operating parameters show to the piston-cam roller contact
good resemblance to the measured external • The distributor valve contributes
leakage flow. Thus, the external leakage in largely to the external leakage of the
the hydraulic motor appears to the largest motor.
extent arise from the distributor valve where
the leakage is dependent upon the clearance
between the valve plates. In the simulation REFERENCES
this valve plate clearance has been assumed
constant, independent upon working pressure. 1. Brokhattingen, E.T., Conrad F.,
The resemblance with measured leakage flow Soerensen, P.R., Trostmann E.,
indicates that the gap height in the distributor Experimental Determination of Static
valve is likely to be rather constant under Characteristics of Hydraulic Motors and
varying operating conditions and that the th
Pumps, Proc. 11 IFAC World Congress,
leakage in the third source of external leakage Tallin, Estonia, Vol. 8 pp. 272-280, 1990.
(between piston and piston bore) is small. 2. Wilson, W.E., Performance Criteria for
The friction between piston and cylinder Positive-Displacement Pumps and Fluid
bore has not been taken into account when Motors, ASME Semi-annual Meeting
determining the load on the piston-cam roller 1948, Paper No. 48-SA-14.
contact. The load in the real motor will thus 3. Schlösser, W., Mathematical Model for
be somewhat lower than the load assumed in Displacement pumps and Motors,
the calculations. Hydraulic Power Transmission, April
The estimated film thickness in full film 1961, pp. 252-257 and 269, May 1961,
lubrication is based upon the values of the pp. 324-328.
designed surface roughness. The actual 4. Thoma, J., Mathematical Models and
surface roughness of the piston-cam roller Effective Performance of Hydrostatic
contact will be changed during the running-in Machines and Transmissions, Hydraulic
time. This change of surface roughness will Pneumatic Power, November 1969, pp.
allow a lower value of film thickness for full 642-651.
film lubrication than that assumed in the 5. Hibi A., Ichikawa T., Torque
simulations. Performance of Hydraulic Motor in
Thermo-elastic deformation will change the Whole Operating Condition From Start to
geometries in some way and thereby affect Maximum Speed and Its Mathematical
the losses, but not in a significant way in full Model, 4th Int. Fluid Power Symposium,
film lubrication regime. In mixed and th
April 16th-18 , University of Sheffield,
boundary lubrication regimes it might have a England, 1975, pp.B3-29-B3-38.

B7
6. Dorey, R.E. Modelling of Losses in 15. Hamrock B. J., Fundamentals of fluid
Pumps and Motors, ist Bath International film lubrication, International Editions
Fluid Power Workshop, University of 1994, McGraw-Hill Inc. USA, ISBN 0-
Bath, UK 1988. 07-025956-9.
7. Zarotti, GI., Nevegna, N., Pump
Efficiencies Approximation and
Modelling, 6th Int. Fluid Power
Symposium, Cambridge, UK, April 1981,
paper C4, pp. 145-164.
8. Rydberg, K-E., On performance
optimization and digital control of
hydrostatic drives for vehicle
applications, Linköping Studies in
Science and Technology. Dissertation
No. 99, Linköping, Sweden, 1983.
9. Conrad, F., Trostmann E., Zhang M.,
Automatic Computer Controlled Bi-
Directional Dynamometer Applied for
Identification of Static Performance and
Experimental Modelling of Losses in
Hydraulic Pumps and Motors, Proc. 3rd
Scandinavian Int. Conference on Fluid
Power, Vol. 1, Linköping, 25-26 May
1993.
10. Huhtala K., Modelling of hydrostatic
transmission — steady-state, linear and
non-linear models, Acta Polytechnica
Scandinavica, Mechanical Engineering
1996, Series No. Me 123, Helsinki.
11. Zhu, D., Cheng, H.S., Arai, T., Hamai,
K., A Numerical Analysis for Piston Skirt
in Mixed Lubrication-Part I: Basic
Modeling, Journal of Tribology, 114, pp
553-562, 1992.
12. Akalin, 0., Newaz, G.M., Piston Ring-
Cylinder Bore Friction Modeling in
Mixed Lubrication Regime: Part I-
Analytical Results, Journal of Tribology,
123, pp 211-218, 2001.
13. Crane Co., Flow of Fluids Through
Valves, Fittings and Pipe, Technical
Paper No. 410M, Crane Co., 1999.
14. SOLVIA software package, SOLVIA
Engineering AB, Sweden.

B8
Paper C
WEAR TESTING AND SPECIFICATION OF HYDRAULIC
FLUID IN INDUSTRIAL APPLICATIONS

Håkan Olsson
ITH, The Institute for Applied Hydraulics
hakan.olsson(d,ith.se, SE-891 61, Örnsköldsvik, Sweden
Jan Ukonsaari
Luleå University of Technology, Division of Machine Elements
ianu@mt.luth.se, SE-971 87, Luleå, Sweden

Keywords: wear, piston, hydraulic motor, hydraulic fluids, boundary lubrication

SUMMARY
Hydraulic radial piston motors are often used as driving units in low speed high torque
industrial hydrostatic transmissions running in more or less continuous operation in a wide
range of different applications. The hydraulic motor type studied in this paper, has
encountered increased wear at low running speeds, i.e. below one rpm. It is in the contact
between cylinder bore and piston where an increased wear rate has occurred. It is of interest to
investigate how this wear is affected by the fluid type used, partly to see if there are
commercially available fluids with which could improve wear protection at low speed of the
motor and partly trying to predict fluids that can cause problems in future low speed drive
applications.
The aim of this paper was to investigate the possibilities to simulate this wear in a model test
and study the influence of different types of commercially available hydraulic fluids on wear
protection and friction. A comparative test was made with a different material combination
where also some additional fluids were utilized. For simulating the contact cylinder
bore/piston in the hydraulic motor, tests was performed in a reciprocating test rig where the
contact geometry is of the cylinder-plate type.
From the study it was concluded that:
• The use of water glycols can cause increased wear in hydraulic motors operating in
low speed applications.
• The FZG test alone appears not to be enough as a hydraulic fluid wear test, at least not
in the tribological conditions present in the performed wear test.
• If the results from this study can be translated to the piston/cylinder bore contact in the
hydraulic motor, the oil specification for the motor should be enlarged to include other
wear tests to complete the FZG test.

Cl
1 INTRODUCTION

Hydraulic radial piston motors are often used as driving units in low speed high torque
industrial hydrostatic transmissions running in more or less continuous operation in a wide
range of different applications. The output demands of a heavy-duty industrial hydrostatic
drive change between different drive applications but in almost all cases there is a strong
demand for a long lifetime and high reliability.
The radial piston motor is of cam-lobe design and
has an even number of pistons that operates in
sequences of working pressure and charge pressure
in a reciprocating motion. The piston is operated
via a cam roller in contact with the curved track
cam ring, which is fixed to the motor casing. The
casing is stationary, which means that the cylinder
block and piston assembly are rotating. In motor
mode, the piston assembly moves radially outwards
under working pressure and inwards during idling.
Figure 1 shows a piston moving outwards. The
pressure in the cylinder bore is generating an active
radial force FR. The contact between the cam roller
and cam ring generates a contact normal force FN.
FT is the tangential force between piston skirt and
Figure 1. Force geometry for the cylinder bore. The size of this force time distance to
piston/cylinder contact.
centre (lever distance) determines the torque
generated by the motor. Consequently, the contact
pressure between piston skirt and cylinder bore is directly proportional to FT. The contact
pressure and the sliding speed of the piston will vary periodically due to the geometry of the
cam ring. The following data holds for the piston/cylinder bore contact:
Piston material: Grey cast iron, Ra = 0.4 gm
Cylinder material: Nodular cast iron, Ra = 0.4 gm
Maximum contact pressure between piston skirt and cylinder bore: 70.5 MPa (at 30 MPa
working pressure)
Maximum sliding speed: 0.01 m/s (at 1 rpm)
Piston stroke: 15.2 mm
At a rotational speed of 1 rpm wear on the piston has been observed in the motors, both in
laboratory tests and in field tests. It is in the contact between cylinder bore and piston where
an increased wear rate has occurred. The location of the wear is at the skirt of the piston
where the tangential force FT is transmitted, see Figure 1. At a speed of 2 rpm and similar
working conditions the wear was substantially lower and above 5 rpm the wear was more or
less negligible. These observations were made using a fully formulated mineral oil with 40
2
MM /s kinematic viscosity. The motor casing is filled with oil, which means that the outer
side of the piston is always exposed to fluid. The low speed of the piston and the lack of

C2
hydrodynamic fluid film forming mechanisms in the contact indicate that these reductions in
wear can be due to squeeze film formation in the contact. Simplified calculations based on
formulas valid for journal bearings with squeeze film action in Hamrock [1] support this
assumption. At increased speed the time required to squeeze the fluid out of the gap between
cylinder and piston appears to be long enough to provide a lubricating film during the piston
stroke.
It is of interest to investigate how the piston wear at speeds below one rpm is affected by the
fluid type used, partly to see if there are commercially available fluids with which it would be
possible to improve the low speed performance of the motor and partly trying to predict fluids
that can cause problems in future low speed drive applications. The aim of this paper is thus
to investigate the possibilities to simulate this wear in a model test and study the influence of
different types of commercially available hydraulic fluids on wear protection and friction. A
comparative test was made with a different material combination where also some additional
fluids were utilized.
Tribological testing can be done in numerous ways and when selecting the appropriate test
method attention must be paid to the following factors [2]:
• Time required for the test • Possibility to control the test
• Possibility to monitor the test • Cost for the test
The primary criterion when making a tribological test is that it reproduces the wear
mechanisms that dominate in the simulated field case [3]. Therefore, it is necessary to identify
the prevailing wear mechanism in order to verify the relevance of the chosen test. Several tips
and hints to aid the evaluation process are given in [2].
Often the hydraulic fluids are specified to meet different major hydraulic pump
manufacturers' specifications. An industry standard vane pump test for indicating the wear
characteristics of petroleum and non-petroleum hydraulic fluids employs a Vickers Vi 04C
vane pump. There are at least three national standards based on the use of this pump; ASTM
D-2882, DIN 51389 and BS 5096 (IP281/77) [4]. The Vickers V104C-test implies that the
vane pump is run for a specified time at specified (fixed) operating conditions. The total
weight loss of the vanes and the ring becomes the quantitative value of wear. However, a
drawback with the standardised vane pump test (ASTM D-2882) is that it does not provide
the necessary correlation required for prediction of the lubricating properties in other pump
configurations [5, 6]. Apart from Vickers 104 C there are also other pump tests to assess the
anti-wear performance of hydraulic fluids, e.g. Denison HF-0 and Vickers 35VQ25 [4].
The FZG test according to DIN 51354 is also often specified in the oil specifications for
hydraulic fluids. With FZG gear tester or FZG gear machine the wear on a pair of spur gears
is measured at different load steps. When larger damage occurs, the test is stopped. The result
is stated in terms of the load stage reached when such damage occurs. According to Mäkelä
[3] the possible conditions or characteristics to measure are: extreme-pressure properties, load
carrying capacity, shear stability, wear etc. According to Reichel [7] the FZG test provides a
forecast of wear occurring in axial piston pumps. A similar correlation for vane pumps and
gear pumps has not been shown.

C3
In the oil specifications for the motor under study it is stated that the fluid used shall fulfil
FZG fail stage minimum 11 as described in DIN 51354. No other tests are specified.
Voitik [8] presents a guide that shall aid the tribologist to organize bench test programs to
solve field problems. The TAN (Tribology Aspect Number) code will help in the transition
from the field problem to the bench test. The code consists of four tribological aspects:
contact velocity characteristics, contact area characteristics, contact pressure characteristics
and entry angle characteristics. A specific TAN code can often be translated to one or more
tribological test specimen descriptions. The TAN code for the contact between piston and
cylinder bore can be taken as 2531 [8]. However, no standard test is available.
For simulating the contact cylinder bore/piston in the hydraulic motor, tests have been
performed in a reciprocating test rig where the contact geometry is of the cylinder-plate type.
Several authors have used this method to investigate different lubricants' effect on friction
and wear under boundary lubrication in sliding steel-steel contacts, e.g. [9] and [10].

2 EXPERIMENTAL METHOD AND MATERIALS

The tests were performed using a Plint & Partner High Frequency Friction Machine (TE
77B). The contact geometry was of the cylinder-plate type, see Figure 2. The machine moves
the cylinder in a reciprocating sinusoidal motion relative to the plate. The dimension of the
cylinder was 06 mm and length 6 mm. The material combination coincides with the piston-
cylinder contact in the motor.
Cylinder material: Grey cast iron, Ra = 0.6 gm
Plate material: Nodular cast iron, Ra = 0.2 gm
Initial Hertzian contact pressure between roller and plate: 460 MPa
Maximum sliding speed: 0.074 mis, amplitude: 2.35 mm, test duration: 10 h

Load As can be seen the contact pressure and


sliding speed was higher in the test than in
the real piston/cylinder contact. This
Plate Cylinder accelerated testing was necessary to speed
up the evaluation. Due to the increase of
the wear scar, the contact pressure
decreases during the tests. At the end of the
test the pressure was 15-150 MPa for the
evaluated fluids. Measurements were made
Figure 2. Plint & Partner High Frequency
with the contact area flooded with oil and
Friction Machine.
the oil bath held at a constant temperature
(40°C), with a maximum error of ± 0.5°C. The wear scar on the cylinder was more
pronounced than the wear scar on the plate and therefore chosen for analysis. The wear scar
on the cylinder from each test was measured and the volume of material removed was
calculated. The wear coefficient k was calculated as the wear volume W divided by the
normal load FN and sliding distance s:

C4
k= (1)
FN • s

The minimum film thickness hmin was calculated for a line contact between cylinder and plate
using the theory of elasto-hydrodynamic lubrication in rectangular contacts according to
Hamrock [1]. The minimum film thickness is given by:

( 10.694 0.568R 0.434


\-4.128 ‘1-10U
min = 1714(r)"°2 ( wiz (2)

E' is equivalent elasticity modulus, w', dimensionless load, 110 absolute viscosity, ri speed
parameter, pressure-viscosity index and 12„ equivalent radius of the sliding surfaces. The
dimensionless film thickness parameter A is given by It4 surface roughness (rms) and the
minimum film thickness. A indicates the state of lubrication:

A— (3)
R2 D2
V id
q,cyiner

A< 1: Boundary lubrication


l<A<3: Mixed film lubrication
3 <A: Full film lubrication
The dimensionless film thickness parameter was calculated for all fluids and it could be seen
that the film parameter was well below 1 for all cases, i.e. boundary lubrication prevailed.

Data for the investigated, fully formulated hydraulic fluids can be seen in Table 1. The
lubricant bases are two mineral oils (Mini), three synthetic esters (SE2), one polyalphaolefin
(PA03), one rape seed oil (VEG4) and one fire resistant water-glycol (FR5). All fluids except
FR and SE 46B fulfil fail stage 11 in the FZG test according to DIN 51354.

Table 1. Fluid data.


Lubricant Kin. visc. 40°C Kin. vise. 100°C Density 15°C
[trun2/s] [mm2/s] [kg/m3]
Min 461 46 6.8 879
Min 68A1 68 8.9 880
SE 462 46 8.6 947
SE 46B2 46 9.2 924
SE 322 36 7.6 916
PAO 463 46 11.6 829
VEG 464 41 9.1 928
FR5 39 - 1080

A comparative study was made in order to state the wear protection abilities with another
material combination. Testing with different materials can show the importance of carefulness

C5
when interpreting results from a test with a material combination not coinciding with the
application. The tested material was a tin bronze sliding on ball bearing steel. Bronze on
hardened steel is a common material combination in machinery:
Cylinder material: Tin bronze, Ra = 1.4 gm
Plate material: Ball bearing steel, Ra = 0.08 gm
Initial Hertzian contact pressure between roller and plate: 360 MPa
Maximum sliding speed: 0.074 m/s, amplitude: 2.35 mm, test duration: 3h
This material combination was evaluated with the fluids presented in Table 1 (except Min
68A) and three additional fluids presented in Table 2. The fluids SE 46A, SE 46B, SE 46C
and Min 68B are specially adapted to water turbine applications. SE 46A is a pure base fluid.
SE 46B and SE 46C are different formulations of SE 46A.

Table 2. Additiona fluid data.


Lubricant Kin. vise. 40°C Kin. vise. 100°C Density 15°C
[truit2/s] [mm2/s] [kg/m3]
Min 68B1 71.7 8.5 888
SE 46A2 45.9 8.1 922
SE 46C2 50.6 8.9 923

3 RESULTS

Results from grey cast iron on nodular cast iron, tin bronze on ball bearing steel and a SEM
study are presented.

3.1 Wear with grey cast iron on nodular cast iron


The wear coefficients for grey cast iron on nodular cast iron are shown in Figure 3. There are
three different homogenous groups of the mean wear coefficient. A homogenous group form a
group of means of wear coefficients within which there are no statistically significant
differences at 95 % confidence level.
Means and 95,0 Percent Confidence Intervals (pooled s) These confidence intervals show that
there is no significant difference
Wear coefficient (mm"3/Nrn)

6,E-9

5,E-9 between the first ranked SE 46B


4,E-9 (6.9E-12) and SE 46 (3.6E-10)
3,E-9
although the mean wear coefficient
2,E-9
differs by a factor of —50. The width
of the confidence intervals depends
1E-9
, -
on the number of tests with each
co
CD CO "e
co Cs1 CD CD
re V'
re lubricant. Fewer tests mean wider
e co 0 w w o
co co w intervals. The synthetic ester SE 46B
g
appears to give the best wear
Figure 3. Wear coefficient of grey cast iron
protection and the fire resistant fluid
on nodular cast iron.
(FR) the worst.

C6
3.1.1 Friction
Means and 95,0 Percent Confidence Intervals (pooled s)
The mean friction values from the tests can
0,14
be seen in Figure 4. The tests were in the
boundary lubrication regime throughout 0,13

their entire duration. SE 46 and SE 46B


0,12
(0.100) seems to have the lowest values of
0,11
friction but not significantly different from
VEG 46 (0.109) at 95 % confidence level. 0,1

0,09
co CV .1 CO Co CD
co
LU
111
("1
83 2
ö'
a_

Figure 4. Friction coefficient of grey cast


iron on nodular cast iron.

3.2 Wear with tin bronze on ball bearing steel


A comparative study was made with tin Means and 95,0 Percent Confidence Intervals (pooled s)

bronze on ball bearing steel. Figure 5 shows


VVear coefficient (mm^3ffilm)

0,000024 -

the mean wear coefficients. The confidence 0,00002-

intervals at 95 % level show which wear 0,000016

coefficients that are significantly different. 0,000012


The synthetic esters show the best wear 0,000008 -
protecting properties and the mineral oil 0,000004
based the worst. 0 -
('a CD CD Co 0 cc < CD 03 CO
CO co LL CD n' co ,r
LU UJ (.2 .cr e 't 0 (0 .G
CO CO W 1.11 Ul LU < sE 2
> til Cl) cri a_ >

Figure 5. Wear coefficient of bronze on


ball bearing steel.

3.3 SEM study

To compare the wear mechanisms operating in the test and in the actual tribo-system a SEM
study was performed. Figure 6 (left) shows the surface of the skirt of a piston that has been
operating for 500 h in a motor running at 1 rpm and 30 MPa working pressure. Figure 6
(right) shows the surface of a test cylinder that has been operating in the test equipment under
the operating conditions specified in section 2.
The material and fluid are the same (grey cast iron, Min 68A) but there is a difference in
sliding distance, 4500 m for the piston and 846 m for the test cylinder. The SEM pictures
indicate that there is some kind of chemical reaction present which is more dominant for the

C7
piston. This can imply higher temperature in the motor, which can cause additives to react. It
can also indicate residues of oil. Abrasive wear scars can be seen on the test cylinder, these
are not present on the piston. This indicates presence of hard particles or parts in the plate
material that can cause these scars. Apart from these differences the pictures show
resemblance indicating that the wear mechanisms are comparable.

mioin 1,et WD
OW, Of 6 Solo) 0 0 kv 50 1000x el- 1 0 ore,

Figure 6. SEM pictures of worn parts. To the left: piston from the hydraulic motor, to the
right: wear scar on a test cylinder.

4 DISCUSSION AND CONCLUSIONS


Synthetic esters appear to give the best wear protection for both material combinations. This,
together with the possibility to make them environmentally adapted, makes them interesting
for the future. SE 46B shows the best wear protection and also the lowest friction in the tests
with grey cast iron on nodular cast iron. This fluid passes FZG fail stage 10.
The fire resistant fluid (FR) shows the worst wear protection of the compared fluids for the
grey cast iron/nodular cast iron contact. Some authors have observed a large wear for water
glycols in steel-steel contacts, [11] and [12]. In [11] the wear of water glycols and petroleum-
based fluids was measured in a four-ball wear tester (FBWT) and in a vane pump test. A
larger wear of water glycols was observed and attributed to the presence of water and an
erosive cavitational attack. According to the authors the additives in these fluids did not seem
to interact effectively with the metal at the surface. One reason for the inferior boundary
lubrication properties of water based fire resistant fluids as compared to mineral oils is,
according to [12], associated with the difficulty of finding boundary additives effective in
water systems.
The use of rape seed oil (VEG 46) resulted in a rather bad wear protection in this test although
the fluid passes FZG fail stage 12. In [9] increased wear was observed for environmentally
adapted hydraulic fluids compared to mineral oil based hydraulic fluids in sliding steel-steel
contacts. This was thought to be a result of the competition between the polar molecules of
the environmentally adapted base oils and the anti-wear (AW) and extreme pressure (EP)
additives in covering the metal surfaces.

C8
One reason of the spread of the wear test results for grey cast iron on nodular cast iron can be
the relatively wide measured variation in hardness, both within and between test pieces.
Another reason could be that the surface roughness showed a rather large variation between
different plates and cylinders.
The SEM study indicated that there is probably not exactly the same wear mechanisms in the
real piston/cylinder bore contact as in the cylinder/plate contact used in the test. However, the
differences are not so big and it is believed that the studied reciprocating wear test can be used
as an initial investigation to predict the boundary lubricating performance of different fluids
in the piston/cylinder bore contact. To verify this, full-scale tests on a hydraulic motor must
be performed.
A different wear ranking was obtained when using grey cast iron on nodular cast iron and tin
bronze on roller bearing steel. This verifies the fact that the complete tribo-system must be
taken into account when selecting an appropriate fluid. For instance, a fluid having good
steel-on-steel anti-wear performance cannot be assumed to have a good steel-on-yellow-metal
anti-wear performance [6].
During the tests with tin bronze on ball bearing steel it could be seen on the friction force
measurements that it was likely that the mixed lubrication regime was entered. The
smoothness of the plate material in combination with the large wear scar developed on the
cylinder can be a reason for this. The probable change in lubrication regime may imply some
irrelevance when making comparisons of the different fluids boundary lubricating properties
in this test.
The conclusions from this study can be summarized as follows:
• The use of water glycols can cause increased wear in hydraulic motors operating in
low speed applications.
• The synthetic ester SE 46B would be interesting to evaluate in an extended hydraulic
motor test.
• The FZG test alone appears not to be enough as a hydraulic fluid wear test, at least not
in the tribological conditions present in the performed wear test.
• If the results from this study can be translated to the piston/cylinder bore contact, the
oil specification for the motor should also include other wear tests to complement the
FZG test.

5 REFERENCES

[1] Hamrock B. J., Fundamentals of Fluid Film Lubrication, International Editions 1994,
McGraw-Hill Inc. USA, ISBN 0-07-025956-9.
[2] Hogmark, S., Jacobsson, S., Hints and Guidelines for Tribotesting and Evaluation,
Lubrication Engineering, No. 5, Vol. 48, pp. 401-409, May 1992.

C9
[3] Mäkelä, K. K., Methods and Apparatuses for Tribological and Rheological Testing,
TRIBOLOGIA-Finnish Journal of Tribology, No.2, Vol. 17, 1998.
[4] Bishop, R. J. Jr., Totten, G.E., Tribological Testing with Hydraulic Pumps: A Review
and Critique, Tribology of Hydraulic Pump Testing, ASTM STP 1310, ISSN 0066-
0558, 1996.
[5] Melief, H. M., Proposed Hydraulic Pump Testing for Hydraulic Fluid Qualification,
Tribology of Hydraulic Pump Testing, ASTM STP 1310, ISSN 0066-0558, 1996.
[6] Young, K. J., Hydraulic Fluid Wear Testing and Development, Tribology of Hydraulic
Pump Testing, ASTM STP 1310, ISSN 0066-0558, 1996.
[7] Reiche!, J., Pump Testing Strategies an Associated Tribological Considerations-Vane
Pump Testing Methods ASTM D 2882, IP281 and DIN 51389, Tribology of Hydraulic
Pump Testing, ASTM STP 1310, ISSN 0066-0558, 1996.
[8] Voitik, R. M., Realizing Bench Test Solutions to Field Tribology Problems by Utilizing
Tribological Aspect Numbers, Tribology: Wear test selection for design and application,
ASTM STP 1199, A. W. Ruff, and Raymond G. Bayer. Eds., American Society for
Testing and Materials, Philadelphia, USA, 1993.
[9] Rieglert, J., Kassfeldt, E., Performance of Environmentally Adapted Fluids at Boundary
Lubrication, Proc. of 23rd Leeds-Lyon Symposium on Tribology, Leeds, UK, September
1996.
[10] Kassfeldt, E., Norrby, T., Rieglert, J., Wear and Friction in Boundary Lubricated
Contacts, Tribology in Environmental Design 2000, ISBN 1-86058-266-4,
Bournemouth, UK, 4-6 September 2000.
[11] Perez, J. M., Hansen, R. C., Klaus, E. E., Comparative Evaluation of Several Hydraulic
Fluids in Operational Equipment, A Full-Scale Pump Stand Test and the Four-Ball
Wear Tester. Part II. Phosphate Esters, Glycols and Mineral Oils, Lubrication
engineering, Vol 46, No 4, May 1990, pp. 249-255.
[12] Ratoi-Salagean, M., Spikes, H. A., The Lubricant Film-Forming Properties of Modern
Fire Resistant Hydraulic Fluids, Tribology of Hydraulic Pump Testing, ASTM STP
1310, ISSN 0066-0558, 1996.

CIO
NR: 2003:15
LULEA ISSN: 1402-1757
TEKNISKA ISRN: LTU-LIC--03/15--SE
UNIVERSI 1

Utbildning
Licentiate thesis
Institution Upplaga
Tillämpad fysik, maskin - och materialteknik 150
Avdelning Datum
Maskinelement 2003-03-11
Titel
Performance Analysis of a Low-Speed High-Torque Hydrostatic Transmission Unit
Författare Språk
Olsson, Håkan Engelska
Sammanfattning
This thesis concerns a study of the performance of an industrial low-speed high-torque
hydrostatic drive system. This type of hydrostatic transmission is commonly used in
continuous operation in a wide range of heavy-duty drive applications. In many
applications the transmissions have to compete with e.g. electromechanical drives, such
as DC or AC electric motors combined with gearboxes. In such situations, energy
efficiency is a key selection criterion in that even a small increase in the efficiency of
high power industrial drives would give substantial savings. Apart from efficiency,
lifetime and reliability requirements are important parameters for industrial drive systems,
as unplanned stops in industrial working processes can be very costly.

The work presented in this thesis is primarily focused on analysing the efficiency
behaviour in the transmission, both on the system level and on the component level.
Attention has also been paid to lifetime issues, with special emphasis on wear occurring
in a sliding contact in a radial piston hydraulic motor.

In Paper A the distribution of power losses in a variable axial piston swash plate pump is
investigated. The pump under study is commonly used in stationary industrial hydrostatic
transmission systems. The churning losses in the pump have been estimated
experimentally by measurements in a test rig. The leakage flow and the power losses in
the contacts between the (cont.)
Granskare/Handledare
Ove Isaksson
URL: http://epubl.luth.se/1402-1757/2003/15
Universitetstryckeriet, Luleå

Вам также может понравиться