Lib Hydraulic
Lib Hydraulic
Lib Hydraulic
Lab AMESim
Hydraulic Library Rev 13
Users guide
How to contact LMS Imagine.Lab
+33 4 77 23 60 30 Phone
+33 4 77 23 60 31 Fax
LMS Imagine
7 place des Minimes Postal address
42300 Roanne - France
Trademarks
AMESim is a registered trademark of LMS Imagine
AMESet is a registered trademark of LMS Imagine
AMERun is a registered trademark of LMS Imagine
AMECustom is a registered trademark of LMS Imagine
LMS Imagine.Lab is a registered trademark of LMS International N.V.
LMS Virtual.Lab Motion is a registered trademark of LMS International N.V.
SysDM is a registered trademark of LMS International N.V.
System Synthesis is a registered trademark of LMS International N.V.
Other product or brand names are trademarks or registered trademarks of their respective holders.
1. Tutorial examples
1.1. Introduction
The AMESim Hydraulic library consists of:
Mechanical library
Used in fluid power application when hydraulic power is translated into mechanical power.
This is a collection of submodels of bends, tee-junctions, elbows etc. It is used typically in low
pressure applications such as cooling and lubrication systems.
You can use more than one fluid in the Hydraulic library. This is important for
modelling combined cooling and lubrication circuits.
The hydraulic library assumes a uniform temperature throughout the system.
If thermal effects are considered to be important, you should use the Thermal
Hydraulic and Thermal Hydraulic Component Design libraries.
There are models of cavitation and air release in the Hydraulic library. Note
also there is a special two-phase flow library. A typical application for this is
air conditioning systems.
In this exercise you will construct the system shown in Figure 1.1. This is perhaps the simplest
possible meaningful hydraulic system. It is built partly from components from the Hydraulic
category (which are normally blue) and partly from the Mechanical category (in green).
The hydraulic section is built from standard symbols used for hydraulic systems. The prime
mover supplies power to the pump, which draws hydraulic fluid from a tank. This fluid is supplied
under pressure to a hydraulic motor, which drives a rotary load. A relief valve opens when the
pressure reaches a certain value. The output from the motor and the relief valve returns to the
tank. The diagram shows three tanks but it is quite likely that a single tank is employed.
The first category contains general hydraulic components. The second contains special
valves.The hydraulic components used in the model you will build can all be found in the first of
these Hydraulic categories. If you click on this category icon, the dialog box shown in Figure 1.2
opens.
First look at the components available in this library. Display the title of components by moving
the pointer over the icons:
Step 1: Use File > New... to produce the following dialog box.
You could also have clicked on the New icon in the tool bar but
if you do this you will have to add the fluid properties icon
yourself.
You get something like Figure 1.4. It is possible that your system may have HL000 associated
with one of the other line runs. These minor variations are dependent on the order in which you
constructed the lines. They will not influence the simulation results.
An important feature to note is that a line run has a special submodel (HL000) which is not a
direct connection. To emphasize this point the line run has a special appearance:
Remember the submodel DIRECT does nothing at all. It is as if the ports at each
end of the line were connected directly together.
In contrast, HL000 computes the net flow into the pipe and uses this to determine the time
derivative of pressure. If the net flow into the pipe is positive, pressure increases with time. If it
is negative, it decreases with time. The pressure created by HL000 is conveyed to the relief
valve inlet. The motor inlet is conveyed by the node and submodel DIRECT.
3. To display the parameters of a line submodel click the left mouse button with the pointer on
the appropriate line run.
Part of the dialog box for HL000 is shown in Figure 1.5. The compressibility of the oil and the
expansion of the pipe or hose with pressure are taken into account together with the pipe
volume. HL000 normally requires the bulk modulus of the hydraulic fluid and pipe wall thick-
ness together with the Youngs modulus of the wall material. From these values an effective
bulk modulus of the combined fluid and pipe walls can be calculated. The effective bulk mod-
ulus of a hose is normally very much less than that of a rigid steel pipe.
5. Click on Close.
Some variables such as a pressure have no direction associated with them. A (gauge) pres-
sure of -0.1 bar indicates that the pressure is below atmospheric. In contrast other variables,
such as flow rate, do have a direction associated with them. A flow rate of -6 L/min indicates
that the flow is in the opposite direction to some agreed standard direction.
Note that you can use the Replay facility to give you a global picture of the results. Figure
1.8 also shows the flow rates in L/min at a time of 10 seconds.
Figure 1.8: Flow rates displayed in replay.
4. To plot a variable associated with a line submodel, click on the corresponding line run.
5. Plot pressure at port 1 for HL000.
Notice how the pressure goes up to just over the relief valve setting (150 bar). During this
time the load speeds up rapidly and actually 'over-speeds'. At this point the motor is demand-
ing more hydraulic flow than the pump can supply. The result is that the pressure must drop
and the relief valve closes. The pressure continues to drop and falls below zero bar gauge.
However, pressure is not like voltage or force. We cannot have a pressure of -100 bar. The
absolute zero of pressure is about -1.013 bar gauge. It is time to introduce two terms.
When pressure falls to very low levels, two things can happen:
All AMESim submodels have hydraulic pressure in bar gauge. The low pressure shown in
Figure 1.10 : Low pressure in the hydraulic pipe is caused by the load speed exceeding its
steady state or equilibrium value. This is highly undesirable behavior as it can result in damage
to the real system.
In reality the starting values we have given for the pipe pressure and load speed are not very
realistic and the prime mover would start from rest or a valve would be used to regulate the flow
to the motor. However, hydrostatic transmission systems like this often do suffer badly from
cavitation and air release problems.
Note that all AMESim submodels display hydraulic volumetric flow rate in L/min. There are two
possible interpretations of this flow rate:
1. There is a very large air content; the pressure drops below the saturation pres-
sure for air in the liquid and air bubbles are formed in the liquid.
2. The pressure drops to the level of the saturated vapor pressure of the liquid and
cavities of vapor form.
3. Extremely high variations in pressure occur such as in certain types of fuel injec-
tion systems.
The first situation is called air release and the second cavitation. If there is cavitation or
significant air release at the inlet to a pump, the flow rate according to the first definition will not
be reduced. However, with the AMESim approach (measuring flow rate at a reference
pressure) it is significantly reduced.
The properties of hydraulic fluids vary a great deal. Modeling them is a very specialist process
This is an example of a component without ports. We cannot connect this icon to any
other.
1. It has an integer parameter index of hydraulic fluid that is in the range 0 to 100 inclusive. This
arrangement means that it is possible to have more than one fluid in an AMESim system.
simplest This has a constant absolute viscosity. The bulk modulus is constant above
the gas saturation pressure and is 1/1000 of this value below the gas saturation pres-
sure. This model is very old but is still used by some AMESim users.It is likely to give
the fastest runs.
elementary This is the default and features a constant liquid bulk modulus with ab-
solute viscosity. The treatment of fluid properties under air release and cavitation is
done.
advanced This gives you access to some cavitation parameters not accessible in the
elementary properties.
advanced using tables This is like the advanced option but you install tables of data
Change the index of hydraulic fluid in FP04-2 to 1. This is a number in the range 0 to 100. If
you look at the other hydraulic components in the system you will find they have index 0 and
3. Set the parameter index of hydraulic fluid to 1. This will change all the parameters in the sys-
tem except FP01 (remember we used Select all and deselected FP01).
4. Specify a batch run in the Run parameters dialog box and initiate the run.
5. Plot several graphs of the batch run to compare results with various air contents: Tools
>Batch Plot .
By zooming in on the curve in regions where the pressure is below 0 bar, you will probably
find some variation in the results, but not to a significant degree.
6. Change the saturation pressure in FP04-2 to 400 bar.
7. Repeat the batch run and update your plot.
Figure 1.17: Pressure in pipe with saturation pressure 400 bar.
The variation between the runs is now very pronounced. The dynamic characteristics of the
system are completely transformed. A few words of explanation are necessary.
Normally the air content of a hydraulic oil is well below 1%. A typical value is 0.1%. It is normally
considered good practice to keep the value as low as reasonably possible. However, in a few
applications, such as lubrication oil in gearboxes, the oil and air are well mixed. In this case,
2.5% is a typical value, and up to about 10% is possible.
A reasonable quantity of air, given time, will completely or partially dissolve in the hydraulic fluid.
The lowest pressure at which all the air is dissolved is called the saturation pressure. For very
slow systems all the air is dissolved above the saturation pressure and partially dissolved below
this pressure. Henrys law gives a reasonable approximation for the fraction of air that is
Some systems are slow enough to stay very close to this equilibrium position (Figure 1.16).
Often classic fluid power systems behave like this. The original saturation pressure is better for
the current example.
However, it does take time for the air to dissolve and this time will not be available in fast acting
systems. Fuel injection systems are a good example of this. Hence with such systems it may be
appropriate to set the saturation pressure artificially high to allow for significant quantities of air
to be undissolved at all pressures.
None of these line submodels takes friction into account. We will suppose that the relief valve is
close to the node but the pump and the motor are at such distances from the node that the
pressure drop along the pipes cannot be ignored. We need to select new pipe submodels that
take friction into account for the pipe runs:
Why did we not choose a more complex submodel that also included inertia? We answer
this question later in this exercise.
3. For the line from the node to the motor, select the submodel HL01.
4. For the line between the node and the relief valve, the submodel DIRECT is already selected
and this is exactly what we want.
Note that ??? indicates that different values are set in the line submodels. Set the index of
hydraulic fluid to 1, diameter of pipe to 10 and pipe length to 5.
2. In FP04-2 reset the saturation pressure (for dissolved air/gas) to 0 bar.
3. Run a simulation with the default run parameters. Do not forget to reset Run Type to Single
run if you have previously run a Batch.
4. Plot the two pressures in HL03.
Figure 1.21: Pressures at the ends of pipe joining pump to node.
Note that there is a large pressure drop along the line. This could be regarded as a sizing
problem but in addition it would be bad practice to site the relief valve so far from the high
pressure point.
This system will enable you to make direct comparisons between results.
3. Go to Run mode and do a simulation. Plot the pressure at the pump outlet (pressure at port
2).
Figure 1.24: Pressure at pump outlet.
We note that the curves are virtually the same. (Try zooming.) There is absolutely no advan-
tage to using HL07 and HL09 instead of HL01 and HL03. If we separated the two systems
It is suggested that:
HL01 should be replaced by HL07 and
HL03 should be replaced by HL09.
In other words with this print interval the lower subsystem is better than the upper. If you re-
plot the pressures at the pump outlets, there are clearly differences. This is what happens if
you zoom.
Figure 1.26: Zoomed pressures at pump outlet.
The violent (and unrealistic) start up has created this oscillation in the pressure of about 56
Hz. It is damped out by 0.1 seconds. Why did we get no warning message in the previous
run? The answer is that a lot of checks are applied to your submodel choices when the run
starts. These take into account the fluid properties, the pipe dimensions and the print inter-
val.
When the print interval was 0.1 seconds, it would have been impossible to see these oscillations
and hence no warnings are issued.
In addition, you force the integrator to compute high frequency phenomena which
are of no interest to you, and which will be invisible at the selected print interval!
The messages on the Warning tab are very helpful. Please read them.
See also 4. Hydraulic Line modeling.
Objectives
Title Value
If you do not change the parameters, the valve will not open. The motor and load
will not move at all! For simplicity leave the other submodel components with their
default settings.
You will not change any parameters but an understanding of the parameters of SV00 will
help you to set those of UD00. The spool has some state variables which are the first two
items in the list. In Parameter mode, their values are the initial values of these state vari-
ables. The spool position is a fraction and so it is a dimensionless quantity in the range -1 to
1.
The next 12 items determine the hydraulic flow characteristics of the valve covering the 4
possible flow paths. When the valve is in one extreme position with fractional spool position
+1, P is connected to A and T to B. In the other extreme position the spool position is -1 and
the connections are A to T and B to P.
When the spool position is 0, there is no flow. To define the flow characteristics of the valve
in the extreme positions a flow rate pressure drop pair is used. The default values of these
are 1 L/min and 1 bar. These values can normally be found in a manufacturer's catalogue.
The parameter critical flow number (laminar -> turbulent) is less important and can be left at
its default value. You can find the details for any submodel if you click on the Help button.
For SV00, this produces the dialog box shown in Figure 1.30.
The rating of the valve rated current is set to 40 mA. This means that an input signal of 40
units will produce a fraction spool position of 1. As the spool moves it behaves like a second
order system. You can specify the natural frequency and damping ratio.
5. Enable Discontinuities Printout in the Run Parameters dialog box.
6. Run a simulation with default run parameters.
7. Select the relief valve component and plot the three quantities:
Flow at relief valve outlet [L/min]
Pressure at relief valve inlet [bar]
Pressure at relief valve outlet [bar]
on the same plot.
Step 2: Plot flow rate against differential pressure for the relief valve.
This is a very common requirement for a 2-port valve, and it involves the use of the Post
processing function.
1. Open the Watch view through the view menu.
2. Select the Relief valve and drag its three variables to the Post processing tab:
4. Plot the new post-processing variable and the flow rate on the same plot:
The system sketch for this exercise is shown in Figure 1.31. The hydraulic actuator (or jack)
moves a load and there is control using position feedback. The position sensor is used to
convert the actuator displacement to a signal. A position duty cycle is specified by a duty cycle
submodel. The duty cycle position is compared with the position indicated by the sensor to
produce an error. The error is subjected to a gain and the signal transferred to the servo-valve.
A further duty cycle supplies an external force to the actuator via the position transducer.
4. When you set the parameters for HJ000, click on the External variables button to call up the
dialog box shown in Figure 1.32.
Figure 1.32: External variables of HJ000
This indicates that a positive velocity means the rod is moving to the right. The greater the
displacement, the further it is to the right. In the current case, a zero displacement and ve-
locity means that the rod and piston are stationary and the piston is at the extreme left end
of the jack.
The meaning of the sign of the acceleration and external force should be clear. A positive
external force opposes the other variables. In other words, it makes a negative contribution
to the acceleration. Hence it is trying to reduce the velocity and displacement.
Remove the dialog box by clicking on Close.
Step 2: Run simulation and plot results
1. Run a simulation setting a final time of 12 s and a print interval of 0.05 s.
2. Plot the following graphs:
Actuator displacement and duty cycle output on the same graph.
Flow rate at the two actuator ports on the same graph.
Flow rate at pump outlet and flow rate at relief valve outlet on the same graph.
Fractional spool position.
The first plot (Figure 1.33) gives an idea of how closely the actual performance matches the
required duty cycle.
3. Plot the output from the summing junction (strictly speaking a differencing junction) that
gives you the position error in m.
4. Try changing the gain attached to the servo valve, the servo valve natural frequency and
damping ratio.
5. Include a high gain value that makes the system unstable.
6. Try introducing a dead band, up to about 10%.
Figure 1.34: Pump and relief valve flow rates.
A typical plot for the flow rates from the pump and relief valve outlets is shown in Figure 1.34.
If you had chosen the pump inlet flow rate instead of the pump outlet flow rate, negative val-
ues would have appeared on the graph. This is easily explained if you click on the External
variables button of the Variable List dialog box. For both ports of the pump a positive flow
rate indicates flow out of the pump. It follows that the flow rate at the pump inlet must be neg-
ative.
7. Plot the two flow rates in the actuator HJ000.
For the this submodel, flow rate is an input on both flow ports. This means a positive flow
rate indicates flow into the component. Figure 1.35 shows typical results. Note how different
Analytical analysis.
AMESim standard runs.
Batch runs.
Linear analysis.
The system is shown in Figure 1.36. The hydraulic jack with the two orifices is the damper and
the accumulator is the spring. It is proposed to use this suspension on the cab of a truck. The
load on each suspension strut is 250kg.I
It follows that
PArod = 250g
From this if we want an operating pressure of about 70 bar, the diameter of the rod must be
about 22.3 mm. We will use a rod diameter of 20 mm and a piston diameter of 40 mm.
2. Set the parameters as shown in the following table.
Submodel Parameter title Value
OR0000 (both parameter set for pressure drop pressure drop/flow rate
instances) pair
Problem 2: The accumulator spring with its precharge pressure of 100 bar is taking no part in
this simulation. The only spring involved at the moment is the hydraulic fluid.
Solution to problem 1:
1. In Parameter mode select Settings > Set final values.
This will give reasonable starting values for state variables. You will find that the piston has
dropped slightly from the mid-position.
2. Reset the following parameters:
3. Run a simulation again and check that the system is in equilibrium with the rod in mid-posi-
tion.
Solution to problem 2:
The two parameters we can vary are the precharge pressure and volume of the accumulator.
For the accumulator to work as a spring, the precharge pressure must be lower than the
The volume of fluid in the jack varies according to the position of the piston. This is due to the
rod. The difference between the minimum and maximum oil volume is
A rod x stroke
which is 0.1 L. The accumulator volume should be a bit bigger than this but certainly not 10 L.
1. Do a run and verify that these values do not disturb the equilibrium.
The values should have changed the spring rate but not the equilibrium position. We need
now to investigate the spring rate.
2. Set the following values
Submodel Parameter title Value
The force value of 2500 N pushes down on the suspension with a value corresponding to the
weight of the vehicle. The force of -2500 effectively takes the complete weight off the
suspension. The slow evolution of the force duty cycle ensures that the system is very close to
equilibrium at all times.
The plot of displacement against force (Figure 1.38) shows the non-linear nature of the spring.
It also shows that the suspension does not bottom out but it does top out.
Figure 1.39 shows that maximum pressure is 160 bar and the minimum is about 40 bar which
occurs when the suspension tops out.
We could continue by doing further analytical calculations. Alternatively we could do batch runs
varying the accumulation pre-charge pressure and accumulator volume and the interested
reader could try this.
However, we will end the exercise by considering the damping of the suspension which is mainly
provided by the two orifices. For simplicity we will assume they are of the same characteristics.
Step 2: Set up a batch run varying the diameters of the orifices with the
vehicle subject to a step change in force
The batch run uses orifice diameters of 1 to 6 mm in steps of 0.5 mm. By zooming in on the plot,
it becomes clear that 3 mm gives a reasonable degree of damping.
2.5 0.835
3 0.482
3.5 0.303
4 0.203
4.5 0.143
5 0.104
5.5 0.078
6 0.060
We can see the evolution of these eigenvalues in a root locus plot. The plot below was obtained
by setting the axis scale as follows:
A more refined search between 2.0 and 3.0 mm would be a good idea but 2.5 mm seems
reasonable.
Density has dimensions of [M/L3] and is expressed in kilograms per cubic meter [kg/m3]. As
mentioned previously density is a function of pressure and temperature:
= ( P, T, nature of fluid )
This function can be approximated by the first three terms of a Taylor series:
( P + P, T + T ) = + ------ P + ------ T
P T T P
with
P
B = ------
T
and
1
= --- ------
T P
This equation is the linearized state equation for a liquid. Using the definition of the density, the
two coefficients and B can also be expressed as:
P 1 V
B = V ------ and = --- ------
V T V T P
B is known as the isothermal bulk modulus or, for simplicity, the bulk modulus and is known
as the cubical expansion coefficient.
Since fluid density varies with the applied pressure, this implies that when a given mass of fluid
is submitted to a pressure change, its volume changes. This phenomenon leads to the definition
where is expressed in units Pa -1 (or m2/N). Considering the relation V = M for a closed
hydraulic circuit the mass is constant, and hence:
d ( V ) = 0 and Vd + dV = 0
it follows that
d------ = dV
-------
V
More usually we use the bulk modulus B, also known as the volumetric elasticity modulus:
-
B = ------
------
P
The relationship between and B implies mass conservation. You must RIGOROUSLY
RESPECT this relationship in the calculations. In the modeling and simulation context of fluid
energy systems, disregarding the relationship between and B leads to abnormal evolutions of
pressure in the closed circuit submitted to compression and expansion cycles. This
phenomenon is strongly accentuated if aeration occurs in the circuit (when dissolved air in the
fluid reappears in the form of bubbles). We shall approach this point by examining the
phenomena of aeration and cavitation.
Air can also have adverse consequences on a fluid compressibility. In liquid, air can be present
in two forms: entrapped and dissolved.
Entrapped air
When the return pipe is not submersed in the tank the liquid jet can entrain some air bubbles in
the tank. Another phenomenon that affects the quantity of air in liquid is leakage.
This air stays in the liquid as cavities and can modify fluid
compressibility. In this context we talk about effective bulk
modulus. Figure 2.2 shows the bulk modulus of a diesel fuel at 40
C with 0, 0.01, 0.1, 1, 10% air. The plot is obtained using the
system shown. The model of the diesel fuel properties is based
on accurate experimental measurements and are designed for
use with injection system which are very fast acting. For this reason air is assumed to be
entrained rather than dissolved.
Dissolved air
Air can also be dissolved in a liquid. A certain quantity of air molecules can be part of the
liquid. In this case the dissolved air does not significantly change the fluid properties.
Suppose the fluid is in equilibrium with a certain percentage of dissolved gas (usually air:
nitrogen and oxygen). Lowering the pressure to just above a critical value called the saturation
pressure induces aeration. This is the process where the dissolved gas forms air bubbles in the
liquid until all the dissolved gases or air are free. The exact point where all the dissolved gas has
come out of solution is difficult to pin-point because it depends on the chemical composition and
behavior of the gas. This is a non-symmetrical dynamic process: the growing process does not
have the same dynamics as when air bubbles disappear. In consequence the total quantity of
If the pressure is dropped further and to just above another critical value called the vapor
pressure, the fluid itself starts to vaporize. This corresponds to a liquid phase change. At a
certain point only fluid vapor and gas exist. In liquid systems the term cavitation usually refers
to the formation and collapse of cavities in the liquid even if cavities contain air or liquid vapor.
The development of a cavity is now recognized as being associated with a nucleation center
such as microscopic gas particles, wear or wall asperities. When the liquid is subjected to a
tensile stress, cavities do not form as a result of liquid rupture but are caused by the rapid growth
of these nuclei.
To understand this, think of beer (or champagne if you prefer) in a bottle. When the bottle is
closed you see no air bubbles and the liquid does not look fizzy. The pressure in the bottle is
above the saturation pressure of the gas in the liquid. When you open the bottle suddenly
bubbles appear and so the dissolved gas (molecules of gas held in the liquid) starts to appear
as gas. In fact the liquid is gas saturated and the atmospheric pressure is less than the
saturation pressure of the liquid. This phenomenon is clearly not cavitation but air release
(aeration). Considering nuclei effects, bubbles form only at particular places in your glass:
around the glass (due to small asperities) and round any particles present in the liquid.
Theoretically, if your liquid was perfectly pure and the wall of the system perfectly regular, air
release or cavitation would occur with great difficulty!
The key point about cavitation is that it is a phase change: the liquid changes to vapor. A
Boiling is a phase change at constant pressure and variable temperature and cavitation is a
phase change at constant temperature and variable pressure.
In any system air release starts first and if the pressure decreases further, cavitation may occur.
This means that, sometimes, people talk about cavitation when the real phenomenon is air
release. Both phenomena can lead to destruction of the material or component.
In both cases it is entrained gas that causes the trouble. When cavities encounter high pressure
in the downstream circuit, these bubbles or cavities can be unstable and can collapse
implosively. The pressure developed at collapse can be large enough to cause severe
mechanical damage in the containing vessel. It is well-known that hydraulic pumps and
pipework can be badly damaged by cavitation and air release.
In all classic hydraulic systems, air release and cavitation must be avoided to prevent material
destruction but sometimes it is required, as with injection systems to prepare the spray
formation.
2.3. Viscosity
Figure 2.5: Viscosity
Viscosity is a characteristic of liquids and gases and is manifested in motion through internal
damping. Viscosity results from an exchange of momentum by molecular diffusion between two
layers of fluid with different velocities. In this sense, the viscosity is a fluid property and not a
flow property.
Figure 2.5 shows the relation between shearing constraint and difference of flow velocity
between two layers
The definition of viscosity was first given by Newton. Between two layers of distance dy, the
exerted force between these two layers is given by:
dU ( y )
F = A ---------------
dy
where U(y) is the velocity depending on the radial position y and dU/dy the velocity gradient.
This proportionality expresses the notion of Newtonian fluid and allows the introduction of
defined as the dynamic viscosity or the absolute viscosity. The dimension of is [ML-1T -1] and
the SI unit is kg/m/s or Pa s. The older unit is the Poise, P, which is 0.1 kg/m/s. However, this is
very small and hence the milli Poise, mP, is the common unit which is 10-4 kg/m/s.
The dynamic viscosity is the constant of proportionality between a stress and the intensity of
shearing between two neighboring layers:
dU ( y )
= --------------- and hence = --------------- = shear stress-
--------------------------
dy dU ( y ) shear rate
---------------
dy
However the absolute viscosity is not very often used in fundamental equations. For example
the dynamics of the elementary volume between the two layers is expressed as:
d dU ( y )
A ------ dy = Ady ---------------
dy dt
In other formulas (e.g. Navier Stokes) the ratio between the absolute viscosity and the density
occurs so often that a new parameter called the kinematic viscosity is introduced:
= ---
The dimension of the kinematic viscosity is [L2 T -1] and so the SI unit is the m2/s. The older
unit of kinematic viscosity is the Stoke, St, which is 10-4 m2/s. However, even this is a very small
unit and hence the centistoke cSt is the common unit with 1 cSt = 10-6 m2/s. This parameter is
Note that the viscosity varies significantly with the fluid temperature.
Normally in absence of air release and cavitation the variation with pressure is not great unless
the pressure is very extreme.
Another important aspect of the viscosity is its influence on the flow conditions of the fluid. We
can distinguish two types of flow conditions:
Laminar flow for which the flow lines are parallel and shearing forces create a pres-
sure drop.
Turbulent flow for which the fluid particles have a disordered, random movement
where
For non-circular cross sections, the hydraulic diameter can be used to determine the Reynolds
number. Hydraulic diameter is defined as follows:
d h = 4xcross sectional area-
----------------------------------------------------
wet perimeter
Orifices (also called restrictions) can be fixed or variable and occur in huge numbers in fluid
systems. Not surprisingly in Engineering courses a mathematical description is presented. This
is usually based on Bernoullis equation and leads to the form
2 ( Pup P down )
Q = C q A -------------------------------------
-
typically 0.7 or
varying with orifice geometry and Reynolds number.
Clearly the flow is laminar for sufficiently small pressure drops which means that Cq is certainly
not constant. One solution is to perform detailed experiments and compute Cq against Reynolds
number. In the context of the orifice (not necessarily circular) the Reynolds number is
Ud
Re = ---------h- where U is a mean velocity and dh the hydraulic diameter. If we take U=Q/A, we end
up with the form Cq =f(Q) and ultimately with
Q = F( Q)
It is possible to work with an implicit relationship like this but we would prefer an explicit formula.
This is provided by introducing another dimensionless number known as the flow number and
denoted by [reference 1]. This is defined as
d 2 ( P up P down )
= ----h- -------------------------------------
-
Both Pup and Pdown are needed. P is not enough because a pressure
drop of 1 bar to 0 bar is not the same as 1001 bar to 1000 bar.
It is not clear which pressure should be used to calculate and .
Possibilities are Pup, Pdown, (Pup + Pdown)/2 or the pressure at the vena
contracta. AMESim uses (Pup + Pdown)/2.
Tables of C q C q ( ) can also be compiled using CFD (computational
fluid dynamics) software.
For high values of , Cq is approximately constant.
The lowest value of at which Cq is approximately constant is called
the critical flow number crit .
The critical flow number for a thin or sharp edge orifice is about 100
and for a long orifice is about 3000.
For a long edge orifice the constant Cq value is also the maximum val-
ue.
For a sharp edge orifice the maximum Cq value can be slightly greater
than the constant value and occurs at a value slightly below crit .
Frictional drag
Submodels belonging to this category are used to model resistance to flow in straight tubes and
conduits. The pressure losses along a straight tube of constant cross-sectional area are
calculated from the Darcy-Weisbach equation:
2
l Q
p = ---- --------------
D 2A 2
min
where:
In straight tubes, the resistance to the motion of a liquid or a gas under conditions of laminar
flow is due to the force of internal friction. This happens when one layer of the liquid (or gas)
has a relative motion compared to the others. These viscosity forces are proportional to the first
power of the flow velocity. We then have:
( Re )
As the Reynolds number increases, the inertia forces, which are proportional to the velocity
squared, begin to dominate. As flow becomes turbulent, there is a significant increase in the
resistance to motion. Part of this increase is due to the roughness of the wall surface. Therefore,
we have:
(Re,rr)
The relative roughness is calculated as the ratio of the average height of asperities to the tube
diameter. See details in Figure 2.9 : Relative roughness:
where:
Dh
is the hydraulic diameter of the pipe.
Commercial steel 45 m
Wrought iron 45 m
Concrete 0.3 to 3 mm
The dependence of the friction coefficient on the Reynolds number and the relative
roughness as shown in Figure 2.10 is often known as the harp of Nikuradse.
Figure 2.10: Evolution of the frictional drag factor with the Reynolds number and with
All lines with friction in the hydraulic category use such a frictional drag factor.
References
[1] McCloy D, Discharge Characteristics of Servo Valve Orifices, 1968 Fluid International
Conference, pp 43-50.
[2] R.C. Binder, Fluid Mechanics. 3rd Edition, 3rd Printing. Prentice-Hall, Inc., Englewood
Cliffs, NJ. 1956.
You can use each icon to install an index of hydraulic fluid in the range 0 to 100. These icons
give you access to a number of submodels which will now be described.
We do not describe the FPDROP submodel here as it is considered obsolete. It is only available
for compatibility with older systems (4.0 and earlier).
FP04
elementary
This option provides the following parameters:
index of hydraulic fluid
density
bulk modulus
absolute viscosity
saturation pressure (for dissolved air/gas)
air/gas content
temperature
polytropic index for air/gas/vapor content
absolute viscosity of air/gas
name of fluid
This option makes the following assumptions:
1. The bulk modulus of the liquid with zero air/gas content is constant. This means
the corresponding density varies exponentially with pressure.
2. The viscosity of the liquid with zero air/gas content is constant.
There is an air release and cavitation model included. Note that name of fluid is a text string
(e.g. cooling water) that identifies the fluid.
When the pressure reaches the saturation pressure of the fluid, some air/gas is released. If the
pressure continues to decrease, the high saturated vapor pressure of the fluid can be reached
and some vapor appears (cavitation: the liquid starts to boil). Remember the fluids used in real
engineering systems are not chemically pure substances. For this reason cavitation is assumed
to occur over a range of pressures and the low saturation vapor pressure is the pressure at
which it is assumed that all liquid has become vapor. All these changes of state strongly modify
the fluid characteristics.
With the elementary option this behavior is taken into account with some reasonable fixed
cavitation parameters. However, with advanced, you can set these values yourself. They are:
Three samples of such files are supplied on the AMESim installation CD:
tblprop1.txt
tblprop2.txt
tblprop3.txt
You can copy these files from the directory:
In mode 1, density and bulk modulus are defined from a reference density, a refer-
ence pressure and a set of tables of bulk modulus values against pressure. Each ta-
ble is written for a given temperature (see tblprop1.txt).
In mode 2, density and bulk modulus are defined from a set of tables of density val-
ues against pressure. Each table is written for a given temperature (see tblprop2.txt).
In mode 3, density and bulk modulus are defined from a reference density, a refer-
ence pressure and a set of tables of speed of sound values against pressure. Each
table is written for a given temperature (see tblprop3.txt).
The viscosity of the fluid is also given in these files after the definition of the density and the
bulk modulus. Two modes are available for the viscosity:
In mode 1 the absolute viscosity is defined from tables of absolute viscosities in cP.
Each table is written for a given temperature (see tblprop1.txt).
In mode 2 the absolute viscosity is defined from tables of kinematic viscosities in cSt.
Each table is written for a given temperature (see tblprop2.txt).
The best strategy if you want to use this facility is to copy these files to a suitable local area and
examine them in an editor. Lines beginning with a '#' are comments and these comments give
further information on how the data is arranged. Then you can select the file that uses the modes
you find suitable, and modify it in order to use your own data.
fuel type
index of hydraulic fluid
It is assumed that these fluids are used in fast acting injection systems and there is no time for
the air content to dissolve or undissolve. The user sets a fixed temperature and the local
temperature is computed using an approximate relationship for an adiabatic change.
simplest
This option gives the simplest hydraulic fluid properties. Its parameters are:
index of hydraulic fluid
density
bulk modulus
absolute viscosity
saturation pressure
air/gas content
temperature
polytropic index for air/gas content
name of fluid
This submodel can be useful in difficult cases. The integrator has an easier task during
cavitation and air release and so it may be possible to get a solution when other methods are
unsuccessful.
FP04 is the only submodel available for this icon. Set the index of hydraulic fluid of the hydraulic
submodels to 1. Change the parameters of the pressure input to get a ramp from 0 to 100 bar
in 10 seconds. Change the parameter name of file specifying fluid properties so that it specifies
your own file tblprop1.txt.
Start a simulation and plot the density, the bulk modulus and the viscosity of the FPROP
submodel against the pressure:
Now edit the values of your file tblprop1.txt and rerun the simulation. Note how the properties
change.
Users should also refer to the hydraulic line demo models provided with AMESim.
4.1. Introduction
The submodels for lines are arranged with the simplest submodel at the top and the most
complex at the bottom. Why are there so many line submodels?
The main problem is the complexity of flow in hydraulic pipes and hoses and the level of detail
necessary for different applications. The following features might be important:
To cause a hydraulic fluid to travel along a horizontal pipe we must provide a pressure gradient
to drive the fluid. This is a resistance effect.
The simplest line submodels are DIRECT and HL000 and these can be described as zero-
dimensional.
The DIRECT line submodel assumes that the two ports are very close together and the fluid and
pipe in between contributes nothing.
HL000 considers the capacitance only. The length of the line is too small for
significant resistance. The fluid velocity and the mass are too small to give significant
inertia. The hydraulic chamber submodel HC00 is essentially the same as HL000.
72
These models solve the 1D Navier-Stokes equations for continuity and momentum presented
below:
------ + --------------
( u ) = 0
t x
2
( u ) ( u + P )
-------------- + --------------------------- + friction = 0
t x
where p is the density, u the mean velocity in the pipe, and friction is the friction term.
In the "lumped distributive" case, these equations are reduced to a set of Ordinary Differential
Equations (ODE) to model the line. These equations are then provided to the AMESim solver
and can be integrated over time. This has the advantage of having a homogenous
implementation. It is possible to apply all features from the usual AMESim environment on line
models. For instance, it is possible to perform linear analysis on such type of line models.
However, this has the drawback of introducing a lot of extra state variables to solve for the
AMESim solver. In particular, in the case of lines with inertia and frequency-dependent friction,
if a lot of nodes are introduced, computational times grow non-linearly. Another drawback of
these ODE methods, it is that the frequency content is truncated in such a way that the highest
eigenvalue of the line model will often be excited in a non-physical manner.
The 1D Navier-Stokes equation can also be solved with Partial Differential Equation solving
methods, such as the Lax-Wendroff scheme. This numerical scheme is applied locally inside
the line component or submodel. This means that the line is no longer solved by the ODE/DAE
solver of AMESim.
These numerical methods have time stepping constraints based on the speed of sound, fluid
velocity and cell size. Basically, the time step of the solver must be smaller than the time it takes
for a wave to travel through a cell. This is generally referred to as the the CFL (Courant
Friedrichs Lewy) condition. The Lax-wendroff lines have their own time-stepping, which is
The Lax-Wendroff solver can be found using the lines HLGCENTER, HLG0020D, HLG0021D,
HLG0022D. The physical equations and numerical implementation behind each submodel are
identical, the submodels only differ in causalities and connectivity.
Since the Lax-Wendroff lines are solved by a different solver than the rest of the system, they
do not introduce extra state variables for the rest of the system. Adding a lot of nodes will mainly
decrease the time stepping in the line and create more rendez-vous points but it will not
introduce extra state variables for the ODE/DAE solver. In a general manner, it is possible to
have line models with a high number of nodes (10, 20, 30...) without dramatic cost in CPU time.
Moreover, the frequency-dependent friction that introduces a lot of states in HL040 is already
implemented in the Lax-Wendroff lines for no overhead in CPU time. Finally, one interesting
feature of the Lax-Wendroff lines is the fact that they do not create spurious numerical
oscillations as do the ODE/DAE methods.
However, there are drawbacks in using the Lax-Wendroff lines. Since they use a different solver,
they introduce a cosimulation based on a physical time-stepping. This cosimulation needs
rendez-vous points that are managed through discontinuities generated by the end pieces of
lines. This can introduce some slowdown in the simulation, especially for steady-state. A typical
ODE/DAE solver will be able to adopt very large time steps when closing to steady-state, which
is not possible using Lax-Wendroff solvers, due to the CFL condition. Another drawback is that
the Lax-Wendroff models do not yet provide all the features provided in lumped and distributive
lines. For instance, the Lax-Wendroff lines uses a rigid wall assumption whereas lumped and
distributive lines can account for wall stiffness.
Finally, since the Lax-Wendroff lines use a local solver, it means that they are not seen by the
ODE/DAE solver of AMESim. Therefore, when one wants to perform linear analysis, the
linearization process ignores the CFD1D lines. In the same way, stabilizing runs cannot be
performed on Lax-Wendroff lines.
The lumped/distributive lines and the CFD1D Lax-Wendroff models are complementary models.
Lumped/distributive lines are very well suited for systems such as large actuation systems and
hydraulic jacks. They can also be used to represent some short lines that do not hold in a 1D
assumption, a Lax-Wendroff line needs at least 3 cells, with typical aspect ratio of L/D=6.
The Lax-Wendroff line models can be used to represent 'longer' lines in which it is necessary to
represent wave propagation accurately. Typically, CFD1D lines can be used to model common
rails in injection systems and connecting pipes between rails and injectors.
72
One approach is to consider the frequency range to be modeled. The user needs to analyze the
system to be modeled, and define a frequency range of interest. Then a simple computation of
the theoretical modes of the line can provide guidelines for the number of nodes necessary to
cover the frequency range of interest (both for Lax-Wendroff models or for ODE/DAE models).
Linear analysis can then be used in AMESim to confirm that the frequency range required is
covered. This approach requires more expertise and more understanding of simulation and
modeling but this is generally the approach recommended. The user can refer to the training
HYD2_SYS that provides in depth information about this method.
Another approach is to follow the guidelines that are presented in the next pages, which give
reasonable models for most cases.
When modeling a line, the user should remember that it is necessary to consider the overall line
to be modeled. A hydraulic line can be cut into 3 different submodels as shown in the following
figure.
However, the modeling should take into account the fact that the physical line is in fact
30+40+60 = 130 mm long. The process illustrated in the next pages takes this into account and
should not be applied independently on each subsection of the line.
AMESim line submodels normally occur in groups of three. The reason for this is the input and
output characteristics of external variables of a submodel. If we connect a pipe to a component,
the component submodel normally does one of the following two things at the connection port:
pressure pressure
Aspect ratio
The checking algorithm in AMESim issues warning messages when you use a one-dimensional
submodel that has an aspect ratio length/diameter ratio less than 6.
Short fat pipes require different submodels than long thin pipes.
For distributed line submodels the line is divided into a collection of cells and the test is that the
cell length/diameter ratio must not be greater than 6.
Dissipation number
If the dissipation number is significantly less than 1, it may be important to consider wave
effects. This motivates the following table.
The result of this test must be qualified by considering the next important number.
Print interval
If this time is significantly less than the print interval, you will never see the waves in plots and
so it is not useful to use a wave dynamics submodel. This is why changing the print interval leads
to the appearance/disappearance of warning messages.
The hydraulic volume submodels HC00 and HC01, which are basically the same as HL000, are
included for completeness. Similarly the zero volume submodel ZEROHV is also included.
Hydraulic Chambers
Zero volume
We are now going to display charts which help decide which line submodel to select. These
must be studied bearing in mind the following notes.
The decision process employed in the charts that follow is very similar to one
employed by AMESim when it checks the suitability of your submodels. If the
submodel is regarded as unsuitable, a warning message is issued, values of
aspect ratio and dissipation number are given for one segment of the chosen
line.
Often the final result from the chart is three submodels such as HL01, HL02,
HL03. Since AMESim will check causality, only one submodel, the one that is
compatible with adjoining submodels, will be offered to you as a choice.
The charts are intended for general guidance and give a good choice most of
the time. However, there are circumstances in which advanced users may
wish to break these rules.
72
where E(x) denotes the largest previous integer of x. In cases where no discretization is
possible, the flowchart may suggest to change the selected line submodel into a sequence of Ni
instances of simpler lines. This number of instances (intentionally limited to 10) is given by:
The following principles are retained for the hydraulic line submodel selection flowchart:
If the user correctly configures a line submodel - even an obsolete one - no warning mes-
sage is displayed.
It sometimes occurs that two different line submodels can be envisaged for the same situa-
tion. If the user has already selected one of them, the other submodel will never be suggest-
ed, unless at least one "wrong" or "perfectible" option is active (e.g. states for frequency-
dependent friction when it is not necessary).
If the user starts a simulation for the first time and takes into account all the warning mes-
sages that were displayed (if any) by performing all the suggested corrections, no warning
message should appear when simulation is launched for the second time.
Each time a discretized line submodel is such that its specified number of nodes is too high,
a warning message suggests reducing it to no more than Nmax to save computation time.
The condition called Cond1 on the flowchart above is given by: Cond1 = (Ac>6) & (Nmax>9) &
(N>5).
A
Advanced . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
Advanced properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
Advanced using tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
Air release . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
Aspect ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
B
Batch parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
Batch run . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
Bernoullis equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
Bulk modulus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
Variation of bulk modulus with pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
C
Cavitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
Cavitation and Air release . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
Cavitation and air release . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
Common parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
Complex line submodels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
Compressibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
Compressibility coefficient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
Cubical expansion coefficient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
D
Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
Diesel
Fluid properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
DIRECT submodel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
Dissipation number . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
Dissolved air . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
Distributed parameter submodels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
Duty cycles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
E
elementary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
Entrapped air . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
F
Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
Laminar, turbulent and transition flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
Flow rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
Fluid compressibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
Fluid properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
FP04 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
Frequency dependent friction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
H
Help
On submodels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
Hydraulic oil
Air content . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
Hydraulic starter system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
I 74
Inertia of fluid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63