Next Article in Journal
The Deformation Characteristics and Lateral Stress of Roadside Crushed Rocks with Different Particles in Non-Pillar Coal Mining
Previous Article in Journal
Financial Development, Clean Energy, and Human Capital: Roadmap towards Sustainable Growth in América Latina
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Properties of Hydraulic Fluid on Pressure Peaks in Axial Piston Pumps’ Chambers

Faculty of Mechanical Engineering and Ship Technology, Gdansk University of Technology, ul. Gabriela Narutowicza 11/12, 80-233 Gdańsk, Poland
Energies 2021, 14(13), 3764; https://doi.org/10.3390/en14133764
Submission received: 17 May 2021 / Revised: 17 June 2021 / Accepted: 18 June 2021 / Published: 23 June 2021
(This article belongs to the Topic Wave and Tidal Energy)

Abstract

:
The article concentrates on the research of the pressure peaks in a working chamber of an axial piston pump. The main focus of the article is the influence of fluid properties on the pressure peaks value. Experimental and analytical methods were used to determine the impact of the fluid density, viscosity, compressibility and aeration. The article describes the pressure peak generation process and explains the theoretical basis of this phenomenon. A model of aerated fluid compression and models of leakage were created using CFD analysis and mathematical equations. These partial models were then merged into one consistent model in Matlab Simulink. The final result was verified by comparing the results obtained from the mathematical model with the results of the experiment. After validation, the model was used to present explicitly how each of the mentioned fluid properties affects the pressure peak value. Additionally, the test stand and measurement equipment used in the experiment were presented along with a short description of the laboratory research and sample results.

1. Introduction

Axial piston pumps are the core of modern hydraulic drives, widely described in the literature [1,2,3]. They are used to supply fluid at high pressures to either complex hydraulic systems, such as the ones mentioned in [4,5], or simpler hydraulic equipment described in [6,7]. Every axial piston pump (Figure 1) consists of a cylinder block (4), moving pistons (3) and is pushed by the swashplate (6). The space created between the piston and the cylinder block is called the working chamber but further in the article it will just be called “the chamber”. Moving pistons suck the fluid into the chamber when they are moving outwards and pump the fluid out of the chamber when moving inwards. In a typical axial piston pump, the commutation plate (7) is used to switch the connection of the chambers between high- and low-pressure channels to ensure the right direction of fluid flow. A change in the connected channels occurs when the volume of the chamber is minimal or maximal. These are the situations when the velocity of the pistons is minimal. Therefore, when the chamber is briefly cut off from both high- and low-pressure channels, the pressure in the chamber does not rise excessively; however, it still causes noise and vibration.
The simplest solution to the problem is not to permit fully cutting off the chamber or to increase the “dead volume” of the chamber. Such solutions are presented in articles focusing on flow ripple reduction [8,9,10], which is an undesired indirect result of pressure peaks. These solutions have their disadvantages; for example, they decrease the volumetric efficiency of the pump. Therefore, it is very important to research and analyze the pressure peak problem in order to find alternative methods to compensate for it.
Pressure peaks have a much bigger impact on the prototype axial piston pump invented at Gdansk University of Technology by Osiecki and Osiecki [11], because it has a different type of commutation unit. The invented pump is called PWK, which stands for an axial piston pump with a cam-driven commutation unit in Polish. The construction of the pump is shown in Figure 2. The work principle of the PWK pump is similar to any other axial piston pump. The pistons (4) are pushed by the swashplates (7). The main difference between the typical axial piston pump and PWK pump is the dual construction. This type of construction demands a different type of commutation unit. In this case, it consists of a commutation tube (9) that switches the connection of the working chamber, marked in yellow, between the high-pressure (red) and low-pressure (blue) channel. The commutation tubes are pushed by the cam (11), which rotates with the shaft (1).
One cycle of the PWK pump’s working chamber volume change at 100% displacement is presented in Figure 3.
Phase I is the beginning of the process when the chamber is disconnected from the high/low-pressure channels, and the pistons’ velocity is near zero. Phase II presents the suction phase when the chamber is connected to the low-pressure channel and the pistons are moving apart. Phase III is the switching phase when the chamber is cut-off and the piston velocity is near zero. Phase IV is the pum** phase when the pistons are getting closer.
Typically, to alter the displacement of an axial piston pump, it is necessary to change the angle of the swashplate. This process and associated problems are described in articles [13,14]. Pumps with a cam-driven commutation unit have a different way of adjusting the displacement. If the cam is rotated separately from the shaft, it changes the moment at which the connection between the chamber and high/low-pressure channels are switched. One cycle of the PWK pump’s working chamber volume change at decreased displacement is presented in Figure 4.
Phase I shows the beginning of the suction process when the fluid is sucked from the high-pressure channel. Phase II presents the switching process when the chamber is cut-off and the pistons are moving apart. In this phase, cavitation may occur. Phase III is the continuation of the suction process when the fluid is sucked from the low-pressure channel. The longer Phase III lasts, in comparison with Phase I, the higher will be the displacement of the pump. Phase IV indicates the moment when the pistons change the direction of their movement. Phase V is the beginning of the pum** process when the pistons are getting closer to each other and the chamber is connected to the low-pressure channel. Phase VI is the moment when high-pressure peaks occur, because the moving pistons still generate a sudden pressure rise in the cut-off chamber. This phase will be the focus of this article and all calculations included in this paper will concern it. Phases VII and VIII show the continuation of the pum** process when the fluid is pumped from the chamber into the high-pressure channel.
The pressure peak p p is defined as the difference between the maximal measured pressure inside the chamber p m a x and the pum** pressure p o , which is described by Equation (1) and presented in Figure 5.
p p = p m a x p o
The pressure peaks are influenced by many factors, which change their values directly or indirectly by affecting the other factors. A graph presenting the grid of connections between the factors and pressure peaks phenomenon is shown in Figure 6.
This article will focus on the impact of the following fluid properties:
  • density;
  • viscosity;
  • compressibility;
  • aeration.
The influence of the operational parameters, such as rotational speed, pum** pressure or displacement setting, is described in detail in other publications [12,15,16].

2. Methods and Models

To assess the effect that fluid properties have on the pressure peaks in working chambers of the PWK pump, two research methods were used: an experiment and a numerical simulation. The experiment provided the reference data used to validate the mathematical model used in the numerical simulation.

2.1. Experiment Description

The experiment was conducted in a laboratory of hydraulics using a 70 kW test stand (Figure 7 and Figure 8). The test stand was equipped with a 2000 L oil tank and temperature stabilization system, which guaranteed stable test conditions. It also allowed changing and maintaining a stable oil temperature within the range from 25 °C to 60 °C.
To measure the value of the pressure peaks, an internal piezoelectric pressure sensor (Figure 9a) was used. It was connected to a working chamber with channels marked with purple color (Figure 9b). The results were recorded with an NI USB-6210 data acquisition (DAQ) module and processed using the “Press-kor” program [12,17]. The obtained results had the form presented in Figure 5. The type and class of measurement equipment used in the experiment was specified in Table 1.
Only the impact of the fluid viscosity could be properly tested experimentally, because it is the only fluid property that can be significantly changed without replacing the fluid inside the tank. The viscosity changes its value with the temperature, when density and aeration remain relatively constant. Temperature’s impact on the bulk modulus is not negligible, but still significantly smaller than on viscosity (Figure 10). Therefore, the test could be described as explicit.
To experimentally verify the influence of the properties other than viscosity, usage of different fluids, such as synthetic or water-based fluids, would be necessary; still, the test would not be explicit, because with the fluid replacement all its properties would be significantly changed.

2.2. Mathematical and Partial CFD Models

To separate and define the influence of particular properties on the pressure peak value, the numerical model of the pressure peak phenomenon was prepared. It was used to simulate the pressure peaks value under conditions that would be hard to measure in the experiment.
The model is based on the definition of the adiabatic, tangent fluid bulk modulus K_A [18]:
K A = V o d p d V A
where:
d p —elementary pressure change in the chamber caused by the elementary volume change d V ;
V o —initial volume of the chamber at the beginning of the compression process.
The selection of the adiabatic bulk modulus was dictated by the rapidness of the process. It is assumed that the adiabatic bulk modulus would be more appropriate to describe the process lasting milliseconds, which is too brief to develop significant heat exchange than the isothermal bulk modulus, which is considered more desirable in long-lasting processes maintaining a constant temperature. The Equation (2) was used in the numerical model where the calculations were made with a very small iteration step. Therefore, it is considered more desirable for the modulus K_A to give more accurate results around a certain point rather than over a wide range. This feature distinguishes the tangent modulus from the secant and therefore it was selected for further calculations.
The Equation (2) was transformed into the following equation:
Δ p i = K A i 1 V k i V o i
where the index i was added to highlight the variables changing iteratively; V o i and V k i are the volumes of the chamber at the beginning and the end of the iteration, respectively; and Δ p i is a pressure change during one iteration.

2.2.1. Fluid Bulk Modulus and Aeration

Fluid bulk modulus is extensively featured in the literature [12,18,19] and there are at least several empiric equations defining its value. For the needs of this article, the Equation (4) [19] describing the secant adiabatic bulk modulus K ¯ A was selected:
K ¯ A i = 1.57 + 0.15 log v 20 10 0.0024 20 τ 10 3 + 5.6 p i M P a
where:
ν 20 —kinematic viscosity in standard conditions (mm2/s);
τ —temperature (°C);
pi—pressure in chamber at the beginning of the iteration (MPa).
To convert secant K ¯ A i to tangent K A i , differential Equation (5) [18] was used:
K A i = K ¯ A i K ¯ A i p i K ¯ A i p i d K ¯ A i d p i
resulting in Equation (6), which was used to calculate the sample hydraulic oil characteristics presented in Figure 11:
K A i = 1.57 + 0.15 log v 20 10 0.0024 20 τ + 3 + 10.2 p i + + 2576 p i 2 1.57 + 0.15 log v 20 10 0.0024 20 τ + 5
Equation (6) does not include aeration of the hydraulic fluid. Depending on the type of fluid, the content of air may vary from 1 to 5%. However, full deaeration of fluid is difficult to achieve. Therefore, it is necessary to include the air content in further calculations. The volumetric fraction of the entrained air is defined by Equation (7) as factor X z :
X z = V G V C + V G
where:
VG—volume of the gas phase in standard conditions;
VC—volume of liquid phase in standard conditions.
Many models describing the value of modulus K_A as a function of factor X_z are presented in the literature [18,19]. These models can be divided into two groups, depending on the definition of the compression process. The first group is based on the ideal gas law, assuming that the air mass in the gas phase remains constant during the compression process. This group includes the following:
  • Merritt model;
  • Nykanen model;
  • Cho and Feldman models (each gives the same result).
The second group is based on the ideal gas law, but also includes the partial or full dissolution of air in the oil during pressure rise. This group includes models such as the
  • Yu model;
  • Ruan–Burton model.
The value of the K_A modulus defined by the models mentioned above is presented in Figure 12. Additionally, for comparison the secant modulus K ¯ _A and tangent modulus K_A were added to Figure 12. Cho’s, Feldman’s and Yu’s models give the same results.
Model Cho was used in further calculations because it showed satisfactory convergence with the research results [20] and has a relatively simple mathematical description with a small number of factors defining it. Model Cho is described by the following equation:
K C h o i = 1 X Z + p n p i 1 k X Z X Z k · p i p n p i 1 k + 1 X Z K A i
where:
k —polytrophic exponent;
p n   —normal pressure (1013.25 hPa).
K C h o i , including aeration, can substitute K A i in Equation (3).

2.2.2. Leakage from the Working Chamber

Equation (3) is still oversimplified, because it does not include any leakage out of the cut-off chamber. Therefore, the volume of leakage V L i was added to the V k i component, formulating the following equation:
Δ p i = K C h o i 1 V k i + Δ V L i V o i
The volume of leakage can be defined as
Δ V L i = Q L i Δ · t i
where:
Q L i is the leakage flow;
Δ t i is the time increment.
There are three main sources of leakage from the cut-off working chamber (11) (Figure 13):
  • Flow through the piston hydrostatic supports, Q h s ;
  • Flow through the annular gaps between the pistons and commutation tube, Q t ;
  • Flow through the gap in the commutation widow, Q o k .
Q L i = Q h s + Q t + Q o k
The first source of the leakage taken into consideration is the hydrostatic support in the slipper. The flow through the support equals the flow Q h s through the spiral choke (Figure 14), which is described in [12,21].
The flow through the choke is almost laminar and given by the following equation:
Q h s = 0.87 Δ p c h · h d l ν · ρ · L d l
where Δ p c h is an unknown pressure drop in spiral choke, and L d l is the length of a screw line. Forces equilibrium Equation (13), formulated based on the dimensions and forces presented in Figure 15, is a starting point for calculations of Δ p c h .
F t = F p h s · cos ξ
Force F t is generated by the pressure p i inside the working chamber and described by the following equation:
F t = p i · π d t 2 4
Force F p h s is generated by the pressure gradient under the slipper and given by the following equation [22,23]:
F p h s = p h s π d p h s z 2 d p h s w 2 8 l n d p h s z d p h s w
where p h s is the pressure in the outlet of the spiral choke. The p h s is calculated using Equation (16):
p h s = p i 2 d t 2 · l n d p h s z d p h s w d p h s z 2 d p h s w 2 cos ξ
which was formulated by transforming Equation (13), in which forces F t and F p h s were substituted with Equations (14) and (15).
Knowing that the pressure drop in the spiral choke is the difference between the pressure inside the working chamber p i and the pressure p h s , Equation (17) may be formulated.
Δ p c h = p i p h s
From Equations (16) and (17), Equation (18) was derived.
Δ p c h = p i 1 2 d t 2 · l n d p h s z d p h s w d p h s z 2 d p h s w 2 cos ξ
Equation (18) can be substituted into Equation (12), which is used to calculate the leakage flow, Q h s .
The second leakage source is the annular gap between the piston and the commutation tube (Figure 2). It is not fully known if the gap is central or skewed. The skewness of the gap may be caused by the situation presented in Figure 16. The compounds of the forces from the swashplates (7), acting perpendicularly on the pistons (4), displace them. Since the pistons (4) are paired and are sliding inside the single commutation tube (9), their skewness cannot be compensated for by the displacement of the tube (9). However, some authors declare that the pistons are centered by the fluid film pressure acting on the pistons inside the tube. Probably, the pistons are only partially skewed thanks to the fluid film pressure and tightly fitted guiding tubes (17) (Figure 2), but not entirely.
Flow through the partially skewed annular gap can be expressed using the following equation [24,25,26]:
Q t i = p i · π · d t · h t 3 12 · ν · ρ · L t · α
where:
α —the skewness coefficient, varying from 1 to 1.38;
h t —average height of the annular gap.
h t = d c t d t 2
The last important location of the leakage from the chamber is the commutation window (Figure 13) Although the commutation bridge is 0.5 mm wider than the commutation window and overlaps it completely, it is known that the leakage flow through the window has the biggest value of all mentioned before. The circular gap around the window is the shortest of all and relatively wide, which causes such as high leakage through it. The simplified shape of the gap is presented in Figure 17.
The gap changes its geometry because the window is moving with the commutation tube. Therefore, the simulation of the flow through the gap was needed. Detailed description of the CFD model can be found in [12]. The results of the simulation are presented in Figure 18. Figure 18a presents the influence of the pressure difference between the inlet and outlet of the gap on the leakage flow. Figure 18b presents how the leakage flow is influenced by a change in dimension bok (Figure 17).
The presented results show the sudden change in flow type. At a gap length bok lower than boko = 0.02 mm, turbulent flow is the dominating type of flow, and when the gap is longer than 0.02 mm, the flow becomes more laminar. The following equation has been formulated to model this phenomenon:
Q o k = Q l a m · f b o k + Q t u r 1 f b o k
where:
Q l a m is the laminar flow component;
Q t u r is the turbulent flow component;
f b o k represents the transition function described by the following logistic function:
f b o k = 1 1 + e 18 b o k b o k r + 1
The laminar component of flow was described as through a gap presented in Figure 19. The shown gap was divided into an infinite number of planar gaps, marked in Figure 19 in red.
The flow d Q l a m through the infinitesimal gap is described by the following equation:
d Q l a m = Δ p · h o k 3 12 · υ · ρ · a d s ^
where:
  • ν—kinematic viscosity;
  • ρ—density;
After mathematical integration, Equation (24) describing Q l a m was obtained:
Q l a m = Δ p · d t u l z · h o k 3 12 · υ · ρ σ o b o k + σ o σ k d σ b o k + r o k r o k 2 d t u l z 2 · sin σ L o k 2 + r o k 2
Since the integral in (24) does not have an analytical solution, it was solved using numerical methods.
The turbulent component of the flow Qtur was calculated using
Δ p = ζ o k · ρ c o k 2 2
where:
ζok—is the local resistance coefficient [12];
cok—is an average flow velocity calculated using
c o k = Q t u r h o k · σ k · d t u l z
Using substitution and transformation, Equation (27) is obtained:
Q t u r = h o k · σ k · d t u l z 2 Δ p ζ o k · ρ
To calculate the coefficient ζok and validate the mathematical model (21), a steady state CFD analysis was conducted in ANSYS CFX. The domain presented in Figure 20a consists of four separate parts:
  • peripheral channel (purple);
  • gap between the commutation tube and the socket (transparent grey);
  • two commutation windows (yellow).
In Figure 20b, the boundary conditions are marked. The inlet (red) boundary condition was the constant pressure inside the working chamber. Constant pressure opening was assigned to the outlet (blue). The peripheral outlet was a simplification, because in the real pump there are two radial channels connected to the peripheral channel. The preliminary tests proved that they do not affect the flow in the analyzed gap; therefore, they were omitted. Two interfaces were created to connect the separate parts of the domain. The orange interface connected the peripheral channel and the gap between the commutation tube and the socket. A green interface connected the gap and commutation window. The grey parts of the model in Figure 20b were no-slip walls. The model allowed to change the gap dimensions, inlet and outlet pressures, and the fluid properties.
The domain was discretized into computational mesh consisting of the hexahedral volumes presented in Figure 21. Each part was divided so that the mesh becomes denser the closer it is to the analyzed gap, which significantly improved the solution and did not extend the computation time excessively.
To simulate the flow in the gap: k-ω, k-ε and Shear Stress Transport (SST) turbulence models were tested; however, the solver was able to converge in all cases only when using the SST model. In some cases, where the solver converged, all models gave similar results. Therefore, the SST model was utilized in further calculations.
A comparison of the results obtained from (21) and the CFD model is presented in Figure 22. The difference between the model and simulation does not exceed 10% if the pressure does not rise over 30 MPa, which is acceptable.

2.3. Discrete Model of the Compression Process

All the above equations were used to prepare a discrete model of the compression process inside the pump’s working chamber (Figure 23).
In this model, time t i and pum** pressure p o were the input parameters. At t o = 0 s the commutation window became fully closed and the compression process began. The compression process was divided into equal timesteps, and for every timestep the actual volume of the working chamber V i was calculated in block V. Additionally, in block ΔV, the elementary change in volume for every timestep was calculated using the following formula:
Δ V i = V i V i 1
The ΔVL block includes Formula (10) and calculates the elementary leakage volume Δ V L i for every timestep. Block K includes Equation (8) and is used to calculate the bulk modulus K i at time t i . Using Equation (9), block K Δ V + Δ V L V calculates the elementary pressure change Δ p i for every timestep. All pressure changes Δ p i are being summed up in block Σ, resulting in an actual pressure peak value p i calculated at time t i . The sum of p i and p o is the actual value of the pressure in the working chamber p r o b i . The pressure p r o b i is used in an algebraic loop to calculate Δ V L i and K i in blocks ΔVL and K. The model was implemented in the Matlab Simulink environment. It allowed to change the geometrical parameters of the pump, the operational parameters, such as pum** pressure or displacement, and of course all the earlier mentioned fluid properties. The results of the simulation using this discreet model are displayed as time domain graphs (Figure 24), presenting the change in pressure in the cut-off working chamber or as a value of a maximal registered pressure.
Since the results presented as series of time domain graphs would be very hard to interpret in the context of fluid properties, in the following part of the article the results will be elaborated upon and presented in another form.

3. Results of the Experiment

Only the impact of viscosity could be measured using the laboratory methods because it is the only parameter that can be significantly changed by changing the fluid temperature, not by changing the fluid itself. The results are presented in Figure 25. It is easy to see that a rising viscosity always increases the pressure peak value, despite the changes in pum** pressure or rotational speed. This means that the leakage flow Q L i in Equation (10) decrease with the viscosity and the pressure inside the working chamber is not unloaded as efficiently as it was, causing a significant pressure peak rise.
Since the laboratory research was very limited and not fully explicit because of the temperature-dependent bulk modulus and density, it was mainly used as a reference for the simulation.

4. Results of the Simulation

The results of the simulation were prepared in Matlab Simulink using the discreet model described in Section 2. A series of numerical calculations were conducted to obtain the time domain graphs similar to the one presented in Figure 24. From the time domain graphs the pressure peak value was read and presented in Figure 26. From it, it is clear that the biggest impact on the pressure peaks are the fluid properties that significantly influence flow resistance: viscosity (Figure 26a) and density (Figure 26b). As it was written before, the hydraulic oil viscosity strongly depends on the temperature, which in normal work conditions can vary from 10 mm 2 s to 250 mm 2 s . It is also possible to use water and water-based liquids as the hydraulic fluid; therefore, the viscosity can drop below 1 mm 2 s . The lower the viscosity is the lower pressure peaks are being predicted, because the increased leakage does not allow to pressurise the liquid in the pump’s chambers.
Density also strongly affects the pressure peaks value; however, to significantly alter the density, it is necessary to change the fluid in the whole hydraulic circuit. The predicted effect of such a change is presented in Figure 26b. The density of the more commonly used fluids can vary from 840 kg m 3 for hot mineral oil to 1261 kg m 3 for glycerine.
What was unexpected, was that the bulk modulus had a relatively low impact on the pressure peaks value. The effect of the bulk modulus change from 1000 MPa to 3000 MPa (Figure 26c) is lower than 10%. Since aeration affects the bulk modulus, its impact on the pressure peaks is even lower (Figure 26d). In normal working conditions, when the fraction of undissolved air does not exceed 3%, the aeration influence is negligible.

5. Discussion and Conclusions

As suggested earlier, the results of the experiment (Figure 25) may not be fully explicit, since during the experiment the change in viscosity was caused by the temperature change. The temperature also affected the bulk modulus and density values (Figure 10). To validate the results of the experiment, a simulation was conducted using models in which all fluid properties were temperature dependent or only viscosity was temperature dependent. The results of this validation are presented in Figure 27. What is easy to notice, is the change in bulk modulus and density caused by the temperature variation did not affect the result significantly. Especially at temperatures below 40 °C, this impact could be described as negligible. Therefore, it is possible to say that the experiment, which was conducted in temperatures varying from 25 °C to 55 °C, presents a true influence of viscosity on the pressure peaks value.
To verify the model (Figure 23), the results of the experiment and simulation are compared in Figure 28. The model gives better results for lower viscosities and the general error does not exceed 10% of the experiment result value, if the fluid temperature is above 40°C. It is also noticeable that the model is much more accurate for a lower rotational speed and pum** pressure. However, even for higher values of rotational speed and pum** pressure, the direction of the pressure peak changes is maintained, and the maximal error of the model does not exceed 30% of the measured value.
In articles, such as [27,28,29], the authors describe the pressure peaks phenomenon in axial piston pumps. However, they concentrate on the design aspects of the problem. They analyzed the impact of the commutation design and the influence of the operational parameters, such as the pum** pressure and rotational speed. Except for fluid viscosity, they did not take into consideration other fluid properties. Nowadays, when a mineral oil is frequently substituted with water- or vegetable oil-based fluids, the impact of the bulk modulus, density or aeration cannot be omitted.
The other group of articles [7,30,31,32,33] shows the consequences of using alternative hydraulic fluids on operational parameters such as overall efficiency, leakage and momentum losses, but they do not include the impact of the fluid change on pressure peaks. Most of them concentrate on the fluid and not on its properties; therefore, it is hard to relate the included data to other non-described fluids.
This article presents a new approach to the problem. It shows how specific parameters of the fluid impact the generation of pressure peaks in axial piston pumps; therefore, it can not only be used as a source of information about commonly used fluids, but also to assess the results of using new, never-tested hydraulic fluids. Additionally, the mathematical models included in the article show the methodology of calculating the compression of the fluid in cut-off chambers of the pump and may be used to assess the leakage in axial piston machines. Future research of the subject will concentrate on pumps with plate commutation and will include pressure-induced deformation of the gaps.

Funding

This research was funded by The National Centre for Research and Development within the framework of program LIDER, grant number LIDER/22/0130/L-8/16/NCBR/2017. Project title: Hydro-mechanical automatic gearbox for agricultural vehicles and heavy machinery. Funding value: 1,197,500.00 PLN.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data available in a publicly accessible repository. The data presented in this study are openly available in [12,34].

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Ivantysyn, J.; Ivantysynova, M.; Ivantysyn, J. Hydrostatic Pumps and Motors: Principles, Design, Performance, Modelling, Analysis, Control and Testing; Tech Books International: New Delhi, India, 2003; ISBN 978-81-85522-16-6. [Google Scholar]
  2. Guo, S.; Chen, J.; Lu, Y.; Wang, Y.; Dong, H. Hydraulic Piston Pump in Civil Aircraft: Current Status, Future Directions and Critical Technologies. Chin. J. Aeronaut. 2020, 33, 16–30. [Google Scholar] [CrossRef]
  3. Kumar, S. CFD Analysis of an Axial Piston Pump; Universitat Politècnica de Catalunya: Barcelona, Spain, 2010. [Google Scholar] [CrossRef]
  4. Cieślik, K.; Krogul, P.; Rubiec, A.; Spadło, K. Concept of a Vehicle for Collecting and Transporting Baled Biomass. In Proceedings of the 10th International Conference on Intelligent Technologies in Logistics and Mechatronics Systems, I℡MS, Panevesys, Lithuania, 21 May 2015; pp. 69–74. [Google Scholar]
  5. Przybysz, M.; Rubiec, A. Simulation Research of Kinematic Discrepancy in Multiaxis Hydrostatic Drive Systems. In Proceedings of the 11th International Conference on Intelligent Technologies in Logistics and Mechatronics Systems, I℡MS 2016, Panevezys, Lithuania, 28–29 April 2016; pp. 131–140. [Google Scholar]
  6. Sliwinski, P. The Basics of Design and Experimental Tests of the Commutation Unit of a Hydraulic Satellite Motor. Arch. Civ. Mech. Eng. 2016, 16. [Google Scholar] [CrossRef]
  7. Śliwiński, P. Influence of Water and Mineral Oil on the Leaks in Satellite Motor Commutation Unit Clearances. Pol. Marit. Res. 2017, 24. [Google Scholar] [CrossRef] [Green Version]
  8. Wu, X.; Chen, C.; Hong, C.; He, Y. Flow Ripple Analysis and Structural Parametric Design of a Piston Pump. J. Mech. Sci. Technol. 2017, 31, 4245–4254. [Google Scholar] [CrossRef]
  9. Xu, B.; Ye, S.; Zhang, J.; Zhang, C. Flow Ripple Reduction of an Axial Piston Pump by a Combination of Cross-Angle and Pressure Relief Grooves: Analysis and Optimization. J. Mech. Sci. Technol. 2016, 30, 2531–2545. [Google Scholar] [CrossRef]
  10. Hong, H.; Zhao, C.; Zhang, B.; Bai, D.; Yang, H. Flow Ripple Reduction of Axial-Piston Pump by Structure Optimizing of Outlet Triangular Dam** Groove. Processes 2020, 8, 1664. [Google Scholar] [CrossRef]
  11. Osiecki, A.; Osiecki, L. Hydrostatic Axial Piston Machine. European Patent 0 742 870, 1999. [Google Scholar]
  12. Patrosz, P. Compensation of Pressure Peaks in Variable Displacement Piston Pump with Cam Driven Commutation, (Original title: Kompensacja Skoków Ciśnienia w Pompie Tłoczkowej z Rozrządem Krzywkowym). Ph.D. Thesis, Gdansk University of Technology, Gdansk, Poland, 2017. Available online: https://mostwiedzy.pl/pl/publication/download/1/kompensacja-skokow-cisnienia-w-pompie-tloczkowej-o-zmiennej-wydajnosci-z-rozrzadem-krzywkowym_32490.pdf (accessed on 10 June 2021). (In Polish).
  13. Załuski, P. Influence of the Position of the Swash Plate Rotation Axis on the Volumetric Efficiency of the Axial Piston Pumps. Sci. Tech. Union Mech. Eng. 2014, 11, 12–15. [Google Scholar]
  14. Załuski, P. Experimental Research of an Axial Piston Pump with Displaced Swash Plate Axis of Rotation. In Advances in Hydraulic and Pneumatic Drives and Control 2020; Stryczek, J., Warzyńska, U., Eds.; Lecture Notes in Mechanical Engineering; Springer International Publishing: Cham, Switzerland, 2021; pp. 135–145. ISBN 978-3-030-59508-1. [Google Scholar]
  15. Osiecki, L.; Patrosz, P.; Zawistowski, T.; Landvogt, B.; Piechna, J.; Żyliński, B. Compensation of Pressure Peaks in PWK-Type Hydraulic Pumps. Key Eng. Mater. 2011, 490, 33–44. [Google Scholar] [CrossRef]
  16. Osiecki, L.; Patrosz, P.; Landvogt, B.; Piechna, J.; Zawistowski, T.; Zylinski, B. Simulation of Fluid Structure Interaction in a Novel Design of High Pressure Axial Piston Hydraulic Pump. Arch. Mech. Eng. 2013, 60, 509–529. [Google Scholar] [CrossRef] [Green Version]
  17. Osiński, P.; Patrosz, P. Badania dynamicznych przebiegów ciśnienia w pompach wyporowych z wykorzystaniem czujników piezoelektrycznych. Napędy Sterow. 2016, 18, 110–114. [Google Scholar]
  18. Gholizadeh, H.; Burton, R.; Schoenau, G. Fluid Bulk Modulus: A Literature Survey. Int. J. Fluid Power 2011, 12, 5–15. [Google Scholar] [CrossRef]
  19. Hodges, P.K.B. Hydraulic Fluids; Arnold, J., Ed.; Wiley: London, UK; Wiley: New York, NY, USA, 1996; ISBN 978-0-340-67652-3. [Google Scholar]
  20. Gholizadeh, H. Modeling and Experimental Evaluation of the Effective Bulk Modulus for a Mixture of Hydraulic Oil and Air. Ph.D. Thesis, University of Saskatchewan, Saskatchewan, SK, Canada, 2013. [Google Scholar]
  21. Zastempowski, B. Badanie Przepływu Oleju Przez Dławik Śrubowy. Ph.D. Thesis, Gdansk University of Technology, Gdansk, Poland, 1981. [Google Scholar]
  22. Złoto, T. Modelowanie Odciążenia Hydrostatycznego i Analiza Zjawisk Przepływowych w Szczelinie Rozrządu Tarczowego Pompy Wielotłoczkowej Osiowej; Monografie/Politechnika Częstochowska; Wydawnictwa Politechn: Częstochowa, Poland, 2007; ISBN 978-83-7193-356-1. [Google Scholar]
  23. Złoto, T.; Kowalski, K. Load of the Slipper-Swash Plate Kinematic Pair of an Axial Piston Pump. MATEC Web Conf. 2018, 157, 08013. [Google Scholar] [CrossRef]
  24. Osiecki, A. Hydrostatyczny Napęd Maszyn; Wydawnictwa Naukowo-Techniczne: Warszawa, Poland, 1998; ISBN 978-83-204-2296-2. [Google Scholar]
  25. Strmčnik, E.; Majdič, F. Comparison of Leakage Level in Water and Oil Hydraulics. Adv. Mech. Eng. 2017, 9, 168781401773772. [Google Scholar] [CrossRef]
  26. Patrosz, P. Influence of Gaps’ Geometry Change on Leakage Flow in Axial Piston Pumps. In Advances in Hydraulic and Pneumatic Drives and Control 2020; Stryczek, J., Warzyńska, U., Eds.; Lecture Notes in Mechanical Engineering; Springer International Publishing: Cham, Swizerland, 2021; pp. 76–89. ISBN 978-3-030-59508-1. [Google Scholar]
  27. Edge, K.A.; Darling, J. Cylinder Pressure Transients in Oil Hydraulic Pumps with Sliding Plate Valves. Proc. Inst. Mech. Eng. Part B Manag. Eng. Manuf. 1986, 200, 45–54. [Google Scholar] [CrossRef]
  28. Ye, S.-G.; Zhang, J.-H.; Xu, B. Noise Reduction of an Axial Piston Pump by Valve Plate Optimization. Chin. J. Mech. Eng. 2018, 31, 57. [Google Scholar] [CrossRef] [Green Version]
  29. Bergada, J.M.; Kumar, S.; Davies, D.L.; Watton, J. A Complete Analysis of Axial Piston Pump Leakage and Output Flow Ripples. Appl. Math. Model. 2012, 36, 1731–1751. [Google Scholar] [CrossRef]
  30. Javalagi, S.; Singireddy, S. Hydraulic Fluid Properties and Its Influence on System Performance; Linko**s Universitet: Linko**s, Sweden, 2012. [Google Scholar]
  31. Rydberg, K.-E. Hydraulic Fluid Properties and Their Impact on Energy Efficiency. In Proceedings of the 13th Scandinavian International Conference on Fluid Power, Linkö**, Sweden, 3–5 June 2013; Linkö** University Electronic Press: Linkö**, Sweden, 2013; pp. 447–453. [Google Scholar]
  32. Knight, G.C. Water Hydraulics. Tribol. Int. 1977, 10, 105–108. [Google Scholar] [CrossRef]
  33. Lim, G.H.; Chua, P.S.K.; He, Y.B. Modern Water Hydraulics—the New Energy-Transmission Technology in Fluid Power. Appl. Energy 2003, 76, 239–246. [Google Scholar] [CrossRef]
  34. Patrosz, P.; Osiecki, L. CFD Analysis of a Flow in Commutation Window in PWK pump [Data Set]. Gdańsk University of Technology. 2021. Available online: https://mostwiedzy.pl/pl/open-research-data/cfd-analysis-of-a-flow-in-commutation-window-in-pwk-pump,530084316516799-0 (accessed on 21 June 2021). [CrossRef]
Figure 1. Axial piston pump with plate commutation unit [12]: (a) cross-section of the pump; (b) commutation plate; P—high-pressure pum** channel; S—low-pressure suction channel; 1—manifold; 2—shaft; 3—piston; 4—cylinder block; 5—slipper; 6—swashplate; 7—commutation plate.
Figure 1. Axial piston pump with plate commutation unit [12]: (a) cross-section of the pump; (b) commutation plate; P—high-pressure pum** channel; S—low-pressure suction channel; 1—manifold; 2—shaft; 3—piston; 4—cylinder block; 5—slipper; 6—swashplate; 7—commutation plate.
Energies 14 03764 g001
Figure 2. Axial piston pump with cam-driven commutation unit (PWK) [12]: P—high-pressure pum** channel; S—low-pressure suction channel; L—leakage channel; 1—shaft; 2—body; 3—socket; 4—piston; 5—spiral choke; 6—slipper; 7—swash-plate; 8—separator plate; 9—commutation tube; 10—cam slipper; 11—cam; 12—adjustment shaft; 13—gearbox; 14—adjustment motor; 15—compensation tube; 16—cover; 17—guiding tube.
Figure 2. Axial piston pump with cam-driven commutation unit (PWK) [12]: P—high-pressure pum** channel; S—low-pressure suction channel; L—leakage channel; 1—shaft; 2—body; 3—socket; 4—piston; 5—spiral choke; 6—slipper; 7—swash-plate; 8—separator plate; 9—commutation tube; 10—cam slipper; 11—cam; 12—adjustment shaft; 13—gearbox; 14—adjustment motor; 15—compensation tube; 16—cover; 17—guiding tube.
Energies 14 03764 g002
Figure 3. Cycle of the PWK pump’s working chamber volume change at 100% displacement [12].
Figure 3. Cycle of the PWK pump’s working chamber volume change at 100% displacement [12].
Energies 14 03764 g003
Figure 4. Cycle of the PWK pump’s working chamber volume change at decreased displacement [12].
Figure 4. Cycle of the PWK pump’s working chamber volume change at decreased displacement [12].
Energies 14 03764 g004
Figure 5. Sample characteristics of the pressure changes inside the working chamber and pressure peak definition [12].
Figure 5. Sample characteristics of the pressure changes inside the working chamber and pressure peak definition [12].
Energies 14 03764 g005
Figure 6. Factors affecting the value of the pressure peaks in the PWK pump.
Figure 6. Factors affecting the value of the pressure peaks in the PWK pump.
Energies 14 03764 g006
Figure 7. Hydraulic test stand at Gdansk University of Technology [12]: (a) driveline and pressure measurement equipment; (b) tank; (c) pump and mechanical parameter measurement equipment; (d) coolers and heaters.
Figure 7. Hydraulic test stand at Gdansk University of Technology [12]: (a) driveline and pressure measurement equipment; (b) tank; (c) pump and mechanical parameter measurement equipment; (d) coolers and heaters.
Energies 14 03764 g007
Figure 8. Test stand hydraulic scheme: Pbad—tested pump; Pobieg—preliminary pump; Pwir—auxiliary pump; SE—electric motors; ZP—pressure relief valves; ZG—main tank; RTemp—heat exchanger; F—filters; np—tachometer; Mp—torque meter; p—manometers; CP—internal pressure sensor; T—temperature sensor; Q—flow meter.
Figure 8. Test stand hydraulic scheme: Pbad—tested pump; Pobieg—preliminary pump; Pwir—auxiliary pump; SE—electric motors; ZP—pressure relief valves; ZG—main tank; RTemp—heat exchanger; F—filters; np—tachometer; Mp—torque meter; p—manometers; CP—internal pressure sensor; T—temperature sensor; Q—flow meter.
Energies 14 03764 g008
Figure 9. Piezoelectric pressure sensor and its localization inside the pump: 2—body; 3—socket; 4—piston; 9—commutation tube; CP—internal pressure sensor. (a) piezoelectric pressure sensor (b) pressure sensor localization.
Figure 9. Piezoelectric pressure sensor and its localization inside the pump: 2—body; 3—socket; 4—piston; 9—commutation tube; CP—internal pressure sensor. (a) piezoelectric pressure sensor (b) pressure sensor localization.
Energies 14 03764 g009
Figure 10. Influence of fluid (mineral oil) temperature on its properties: ΔK-bulk modulus change; Δν-kinematic viscosity change; Δρ-density change (reference values measured at normal conditions).
Figure 10. Influence of fluid (mineral oil) temperature on its properties: ΔK-bulk modulus change; Δν-kinematic viscosity change; Δρ-density change (reference values measured at normal conditions).
Energies 14 03764 g010
Figure 11. Characteristics of the adiabatic, tangent bulk modulus K A as a function of pressure p and temperature τ for Total Azolla ZS 46 oil [12].
Figure 11. Characteristics of the adiabatic, tangent bulk modulus K A as a function of pressure p and temperature τ for Total Azolla ZS 46 oil [12].
Energies 14 03764 g011
Figure 12. Comparison of the bulk modulus models for Total Azolla ZS 46 oil.
Figure 12. Comparison of the bulk modulus models for Total Azolla ZS 46 oil.
Energies 14 03764 g012
Figure 13. Localization of the leakage sources [12]: red—flow through the hydrostatic supports; purple—flow through the annular gaps between the pistons and commutation tube; green—flow through the gap in the commutation widow; 2—body; 3—socket; 4—piston; 5—spiral choke; 6—slipper; 7—swash-plate; 8—separator plate; 9—commutation tube; 17—guiding tube.
Figure 13. Localization of the leakage sources [12]: red—flow through the hydrostatic supports; purple—flow through the annular gaps between the pistons and commutation tube; green—flow through the gap in the commutation widow; 2—body; 3—socket; 4—piston; 5—spiral choke; 6—slipper; 7—swash-plate; 8—separator plate; 9—commutation tube; 17—guiding tube.
Energies 14 03764 g013
Figure 14. Spiral choke [5].
Figure 14. Spiral choke [5].
Energies 14 03764 g014
Figure 15. Load distribution on the piston and slipper [5].
Figure 15. Load distribution on the piston and slipper [5].
Energies 14 03764 g015
Figure 16. Skewed annular gap between the pistons (4) and commutation tube (9).
Figure 16. Skewed annular gap between the pistons (4) and commutation tube (9).
Energies 14 03764 g016
Figure 17. Simplified geometry of a gap in the commutation window: red—inlet; blue—outlet.
Figure 17. Simplified geometry of a gap in the commutation window: red—inlet; blue—outlet.
Energies 14 03764 g017
Figure 18. Influence of the pressure difference between the inlet and outlet of the gap and the gap length on the leakage flow: (a) characteristic of the flow as a function of pressure difference Δp for different gap lengths bok; (b) characteristic of the flow as a function of gap length bok for different pressure differences Δp.
Figure 18. Influence of the pressure difference between the inlet and outlet of the gap and the gap length on the leakage flow: (a) characteristic of the flow as a function of pressure difference Δp for different gap lengths bok; (b) characteristic of the flow as a function of gap length bok for different pressure differences Δp.
Energies 14 03764 g018
Figure 19. Graphical representation of the model of the commutation window gap.
Figure 19. Graphical representation of the model of the commutation window gap.
Energies 14 03764 g019
Figure 20. Domain geometry and border conditions. (a) Parts of domain: purple—peripheral channel; transparent grey—gap between the commutation tube and the socket; yellow—commutation windows. (b) Border conditions: red—inlet; blue—outlet; green and orange—interfaces.
Figure 20. Domain geometry and border conditions. (a) Parts of domain: purple—peripheral channel; transparent grey—gap between the commutation tube and the socket; yellow—commutation windows. (b) Border conditions: red—inlet; blue—outlet; green and orange—interfaces.
Energies 14 03764 g020
Figure 21. Discretization of the domain into computational mesh: (a) complete model; (b) commutation windows; (c) gap between the commutation tube and socket [26].
Figure 21. Discretization of the domain into computational mesh: (a) complete model; (b) commutation windows; (c) gap between the commutation tube and socket [26].
Energies 14 03764 g021
Figure 22. Comparison of the characteristics obtained from the CFD simulation and (21): (a) characteristic of the flow as a function of the pressure difference Δp for different gap lengths bok; (b) characteristic of the flow as a function of the gap length bok for different pressure differences Δp.
Figure 22. Comparison of the characteristics obtained from the CFD simulation and (21): (a) characteristic of the flow as a function of the pressure difference Δp for different gap lengths bok; (b) characteristic of the flow as a function of the gap length bok for different pressure differences Δp.
Energies 14 03764 g022
Figure 23. Simplified block diagram of the discrete simulation model of the compression process.
Figure 23. Simplified block diagram of the discrete simulation model of the compression process.
Energies 14 03764 g023
Figure 24. Sample result of the simulation using the Simulink discreet model.
Figure 24. Sample result of the simulation using the Simulink discreet model.
Energies 14 03764 g024
Figure 25. Influence of viscosity on pressure peaks value: (a) measured at a constant pum** pressure of 10 MPa, pump displacement set to 20% of the maximal value and a variable rotational speed; (b) measured at a constant rotational speed, pump displacement set to 20% of the maximal value and a variable pum** pressure.
Figure 25. Influence of viscosity on pressure peaks value: (a) measured at a constant pum** pressure of 10 MPa, pump displacement set to 20% of the maximal value and a variable rotational speed; (b) measured at a constant rotational speed, pump displacement set to 20% of the maximal value and a variable pum** pressure.
Energies 14 03764 g025
Figure 26. Influence of the properties of hydraulic fluid on the pressure peaks value: (a) influence of viscosity; (b) influence of density; (c) influence of the bulk modulus; (d) influence of aeration. All simulations were conducted with the following settings: pum** pressure p_o = 10 MPa, rotational speed n = 1500 rpm, and displacement setting x = 20%. All constant fluid parameters were properties of the Total Azolla 46 oil, calculated at a temperature of 50°C and at 0% aeration.
Figure 26. Influence of the properties of hydraulic fluid on the pressure peaks value: (a) influence of viscosity; (b) influence of density; (c) influence of the bulk modulus; (d) influence of aeration. All simulations were conducted with the following settings: pum** pressure p_o = 10 MPa, rotational speed n = 1500 rpm, and displacement setting x = 20%. All constant fluid parameters were properties of the Total Azolla 46 oil, calculated at a temperature of 50°C and at 0% aeration.
Energies 14 03764 g026aEnergies 14 03764 g026b
Figure 27. Comparison of the pressure peaks values predicted using a model in which only viscosity was temperature dependent and model in which all fluid properties were temperature dependent.
Figure 27. Comparison of the pressure peaks values predicted using a model in which only viscosity was temperature dependent and model in which all fluid properties were temperature dependent.
Energies 14 03764 g027
Figure 28. Comparison of the pressure peak values obtained for variable temperatures (viscosity) from the experiment and simulation: (a) obtained at a constant pum** pressure of 10 MPa, displacement setting of 20% and variable rotational speed; (b) obtained at a constant rotational speed, displacement setting of 20% and variable pum** pressure (experiment results are marked with a thicker solid line, and the simulation results are marked with a thinner dotted line).
Figure 28. Comparison of the pressure peak values obtained for variable temperatures (viscosity) from the experiment and simulation: (a) obtained at a constant pum** pressure of 10 MPa, displacement setting of 20% and variable rotational speed; (b) obtained at a constant rotational speed, displacement setting of 20% and variable pum** pressure (experiment results are marked with a thicker solid line, and the simulation results are marked with a thinner dotted line).
Energies 14 03764 g028
Table 1. Specification of the measurement equipment used in the test stand.
Table 1. Specification of the measurement equipment used in the test stand.
No.Measured ValueMeasuring InstrumentClass of the InstrumentMeasuring RangePrecision/SensitivityMax Uncertainty
1Pressure at suction channelMPS-02 manometer0.22.4 MPa0.001 MPa0.006 MPa
2Pressure at pum** channelHBM manometer0.250 MPa0.05 MPa0.15 MPa
3Pressure at pump’s chamberM105C22 piezoelectric sensorn/a34.47 MPa/4888 mV6.894 kPa/97.43 µV437.8 kPa/59.56 mV
PA-3000 signal amplifiern/a20 Vn/a72.5 mV
NI-6210 DAQ modulen/a−10÷10 V0.306 mV0.402 mV
4Flow ratePT-200 flowmeter0.2200 dm3/min0.01 dm3/min0.41 dm3/min
5Output torqueHBM torque meter0.2500 Nm0.01 Nm1.01 Nm
6Rotational velocity of pump shaftIncremental encodern/a3000 RPM1 RPM1 RPM
7Oil temperatureThermocouple1.070 °C1 °C1.7 °C
8TimeQuartz clock of NI-6210 DAQ modulen/a3 s50 nsno data
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Patrosz, P. Influence of Properties of Hydraulic Fluid on Pressure Peaks in Axial Piston Pumps’ Chambers. Energies 2021, 14, 3764. https://doi.org/10.3390/en14133764

AMA Style

Patrosz P. Influence of Properties of Hydraulic Fluid on Pressure Peaks in Axial Piston Pumps’ Chambers. Energies. 2021; 14(13):3764. https://doi.org/10.3390/en14133764

Chicago/Turabian Style

Patrosz, Piotr. 2021. "Influence of Properties of Hydraulic Fluid on Pressure Peaks in Axial Piston Pumps’ Chambers" Energies 14, no. 13: 3764. https://doi.org/10.3390/en14133764

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop