Applied Computational Fluid Dynamics by Hyoung Woo Oh - HTML preview

PLEASE NOTE: This is an HTML preview only and some elements such as links or page numbers may be incorrect.
Download the book in PDF, ePub, Kindle for a complete version.

maintain as constant as possible the pressure drop across its metering edges. This action is normally influenced by a local pressure compensator which could have either a single or a

double stage configuration, and is usually placed upstream or downstream the control valve

centre. Therefore, for a given operating position of the control valve spool, and for the flowrate across the efflux area of metering orifices the pressure drop can be maintained constant independently by the actuator work-load. Fig. 15 depicts the closed centre load-sensing

proportional control valve studied here. It is designed for operational field limits up to 100

l/min as maximum flow-rate and up to 350 bar as maximum pressure. As shown, the centre

presents a Z connection between the high pressure port (P, in red) and the internal volume

(P1, magenta), which is metered in all spool directions by a twice notched edge. At the same time, both connections between the internal volume and the actuator ports (A and B, green),

and those between the actuators ports and the discharge line ports (T, blue) are metered by

multiple notched edges. It is worth mentioning that the central volume indicated as P1

normally hosts the local pressure compensator (not included in the sketch in Fig. 15). The

Figure presents also a view of the proportional control valve spool including a zoomed view

index-273_1.png

Multidimensional Design of Hydraulic Components and Systems

263

Fig. 15. Main section of the control valve and view of the spool’s metering edges

of the design of the central edge. As shown, two equal notches located at 180° characterize

each side of the central edge, and each couple of notches has a 90° rotation with respect to the other one. Moreover, the notches on the left side of the central edge meter the hydraulic power during the spool motion from the centre to the right (when P1 opens to A, and B

opens to T), while the notches on the right side works during the motion from the centre to

the left (when P1 opens to B, and A opens to T). In this way, the high pressure port P is

connected to the control valve internal volume through an asymmetric path, which depends

on the spool direction of motion. The CFD domain used in the simulations, see Fig. 16,

includes the metering edges of the proportional control valve (upper, lower, front and back

notch), the inlet duct and the outlet region up to the central volume where usually a local

pressure compensator is positioned. The considered domain is marked with a dashed line in

Fig. 15. The corresponding mesh is created using an unstructured grid paying particular

care to the metering areas. Local refinements are used to obtain a large number of cells in

the critical sections and at small opening positions. Moreover, wall cell layers are employed in order to have the proper wall cell height accordingly to the adopted turbulence model.

The average mesh resolution is set to 0.3 mm corresponding to an overall cell number equal

to 3 million elements. Fig. 16 b) shows the mesh of the zoomed views of a the metering edge.

In the numerical analysis, several spool positions are considered in order to simulate

different operating conditions of the control valve. In particular five spool displacements are investigated when the inlet and outlet regions are connected by means of the upper and

lower notches (i.e. direct flow through the notches) and the corresponding five spool

displacements in which the front and back notches connect the inlet and outlet of the valve

(i.e. inverse flow through the notches). The former spool positions are considered as positive while the latter ones are assumed as negative even though the opening area is the same,

since the spool displacement is symmetric with respect to the valve centre. Table 4 details all the operating conditions used in the analysis. As can be noticed from the Reynolds number

listed in Table 4, all the cases are fully turbulent. Therefore, in the simulations carried out for the present work the turbulence is modelled by means of the two zonal version of the k-ω

model, known as the shear stress transport model. Furthermore, careful attention is paid in

order to comply with the requirements of the turbulence model adopted. The results of the

numerical analysis of the load-sensing proportional control valve are discussed in terms of

discharge coefficients, pressure and velocity fields, flow forces and flow acceleration angles.

In particular the effect of the direct and inverse flow through the notches is highlighted. Fig.

17 a) depicts the comparison between the reference and calculated inlet pressures. In the

simulations the outlet pressure is assumed as a constant boundary condition, while the inlet

index-274_1.jpg

264

Applied Computational Fluid Dynamics

Fig. 16. a) The geometry of the control valve included in the CFD domain and b) detail of the mesh in the notch region for a mid range opening position of the spool

Spool

displacement

Reynolds

(x/xmax)

number

-1.00 3573

Inverse flow

-0.78 4209

through the front

-0.65 3289

and back notches

-0.50 2174

-0.35 1422

0.35 1423

Direct flow

0.50 2176

through the upper

0.65 3290

and lower notches

0.78 4210

1.00 3561

Table 4. Spool displacements and operating conditions used in the simulations

pressure is calculated as a consequence of the fluid – dynamics losses through the metering

edges. As a first result, it can be noticed that even though the boundary conditions are

symmetrical with respect to the control valve centre, the predicted inlet pressure is slightly different when comparing the direct and inverse flow. This behaviour is more evident when

comparing the discharge coefficients, see Fig. 17 b). The discrepancy can be attributed to

two main reasons. First, in the direct flow (i.e. positive spool displacements) the effective area decreases progressively as the liquid flows through the notches up to the vena

contracta section close to the notch outlet; conversely, in the inverse flow, the area decreases abruptly at the notch entrance. Second, the front and back notches have very similar

geometrical boundaries both at the inlet and at the outlet, while the geometrical volumes

downstream the upper and lower notches differ significantly. In fact the liquid flowing out

from the upper notches is heading for the outlet boundary, whilst the flow out from the

lower notch finds a confined region causing further pressure losses. These reasons have

opposite effects on the discharge coefficient; in fact the first one would advantage the direct flow while the second reason would advantage the inverse flow. The numerical simulations

show that the trade-off between these two contrasting trends results in a better discharge

coefficient for the inverse flow. This behaviour is also confirmed when comparing the flow

index-275_1.png

index-275_2.png

index-275_3.png

Multidimensional Design of Hydraulic Components and Systems

265

Fig. 17. Comparison a) between reference and calculated pressures at the valve inlet and

outlet and b) between reference and calculated discharge coefficients

fields obtained for the different operating conditions. As an example, Fig. 18 plots the

pressure and velocity fields on a cut section through the symmetry plane of the front and

back notches (inverse flow) and the upper and lower notches (direct flow) for a small

displacement case. While the flow through the front and back notches look very similar, the

pressure and velocity fields are significantly different for the upper notch and the lower

notch. In particular, the flow exiting the lower edge hits the outlet chamber wall and

bounces back into the spool volume and then it is redirected to the outlet boundary. This

Fig. 18. a) Pressure and b) velocity distributions on a cut section through the symmetry

plane of the metering section for the inverse and direct flow (x/xmax=0.35).

Fig. 19. Pressure distributions on the spool chambers used for the calculation of the flow

forces for a) the inverse flow and b) the direct flow)

index-276_1.png

266

Applied Computational Fluid Dynamics

Fig. 20. a) Flow forces exerted on the valve spool in the axial direction and b) ratio between the total axial flow forces in case of direct and inverse flow

more complicated flow pattern increases the fluid – dynamics losses and thus determines a

smaller discharge coefficient. By integrating the pressure distribution over the surface of the spool chambers’ walls, it is also possible to calculate the axial forces exerted on the spool by the flow. Fig. 19 depicts the pressure contours on the walls of the spool chambers for all

notches both in case of direct and inverse flow (as an example only the case with

x/xmax=0.35 is plotted). It can be noticed that the pressure contour is quite uniform for the non operational notches, while the pressure varies significantly throughout the walls of the notches through which the oil flows. This behaviour results in a very low integrated value of the flow induced forces on the non operational chamber of the spool, as can be seen in Fig.

20. When comparing the integrated values of the forces for the direct flow (upper and lower

notches) and the inverse flow (front and back notches), it is possible to outline that the

behaviour is clearly not symmetrical with respect to the control valve centre, see Fig. 20 a).

In fact, even though the trend looks similar and in both cases the sign of the forces reverses moving from the small to the large displacements, the magnitude is rather different. Fig. 20

b) details the ratio between the total axial forces obtained in case of direct and inverse flow.

The forces have the same sign only at mid range displacements and the inverse flow

induced forces are initially lower than the direct flow ones but approaching the maximum

displacement they become much larger. Finally, by using the average velocity vector at the

exit of the four notches it is possible to estimate the flow acceleration angle for each

metering edge. Fig. 21 shows the convention used in this analysis for the calculation of the efflux angle with respect to the coordinate system axes. When values larger than 90 degrees

are found, the negative supplementary angle is considered. Fig. 22 a) shows the efflux

angles for the four metering sections as a function of the spool displacement. The angles are calculated for each notch only in its operational spool opening range. It is interesting to

notice that each couple of notches has a quite similar behaviour. While this result could be expected for the front and back notches due to their similar geometrical boundaries, it is less evident in the case of the direct flow, where the downstream region is very different for the upper and lower notches. Furthermore, when comparing the efflux angles obtained for the

direct and inverse flows, it can be remarked that the trend looks similar. In fact, smaller

angles are characterizing the small spool displacements, as a consequence of the Coanda

effect, whereas they are increasing moving towards the maximum opening. Nevertheless, in

the average the direct flow results in larger flow acceleration angles; this behaviour is likely due to the fact that the main flow direction change happens close to the notch exit, while in the inverse flow the stream bends abruptly just entering the notch. By knowing the flow

index-277_1.png

index-277_2.png

Multidimensional Design of Hydraulic Components and Systems

267

acceleration angles it is then possible to calculate the theoretical values of the axial forces, th

F , through the von Mises theory:

F  2

  C C  cos   

th

d

v

p A

(7)

where Cd and Cv are the discharge coefficient and the velocity coefficient (in the following assumed equal to 0.98), Δ p is the pressure drop across the metering edge, A is its geometrical area and is the efflux angle. Fig. 22 b) depicts the comparison between the theoretical flow force and the one predicted by using the numerical simulation (grey lines). In the CFD

results, the contributions of each metering edge are also plotted (red and black lines). The theoretical calculations and the numerical predictions are significantly different, and in

particular by using the von Mises theory it is not possible to determine the change of

direction moving from low to high spool displacements both for the inverse and direct flow.

Fig. 21. Convention used for the calculation of the efflux angles

Fig. 22. a) Efflux angles for the different notches as a function of the spool displacement and b) comparison between theoretical and predicted axial forces

5. Fully transient modelling of hydraulic components

The numerical analysis of the hydraulic valve addressed in the previous section is further

deepened by adopting a fully transient numerical approach. The main issue in this type of

analysis is the moving mesh methodology employed to simulate the spool displacement

profile versus time. To accomplish this task, the mesh changer libraries of the OpenFOAM

code are modified to account for the linear motion of a portion of the geometry sliding over a second portion. The mesh motion is resolved by using a Generalized Grid Interface (GGI)

index-278_1.jpg

268

Applied Computational Fluid Dynamics

Fig. 23. PWM control applied to the spool motion

approach (Beaudoin & Jasak, 2008), originally developed for turbomachinery applications

and modified to include not only rotational motion of the moving grid but also the linear

displacement of a valve spool. The spool is moved accordingly to the PWM control profile

depicted in Fig. 23. The other models adopted in the analysis are the same as the ones

detailed in the previous sections. The valve is tested under several operating conditions. In particular, three constant flow rates at the inlet are considered, while a forth case is

accounted for, in which the pressure across the valve is held constant. The four cases are

first simulated by using a steady state approach and then a full transient approach. For the steady state simulations, four spool displacements are studied spanning from an almost

closed position to mid range and maximum displacement. In case of steady state analysis

there is no difference between the opening and the closing travels; on the contrary for the

transient approach different inertia phenomena may arise. Table 5 lists the simulated cases, detailing the steady states and transient simulations. The comparison between the results

obtained by the steady state calculations and the fully transient ones are discussed in

terms of total pressure drop across the component, discharge coefficients, pressure and

velocity fields, and flow acceleration angles. In Fig. 24 a) the total pressure drop across the valve is depicted for Cases # 1 to 3 and Cases # 5 to 7. The values predicted by the steady

state calculations and the transient ones do not differ remarkably; particularly in the

transient simulations the pressure variation during the opening and closing travels are

quite similar. A large discrepancy can be seen when considering constant pressure at the

boundaries. Fig. 24 b) plots the flow rate through the valve when holding constant the

pressure drop across the component, i.e. 20 bar. The values predicted by the steady state

simulations at small spool displacements are very close to the ones calculated during the

opening phase with the transient approach. At large spool displacement the saturation

flow rate resulted to be lower for the transient approach than for the steady state

simulation. It is interesting to notice the difference between the flow rate calculated

during the opening and the closure of the spool. The inertia of the flow through the

metering edge of the valve causes a small fluctuation as the spool reverses it direction,

while at mid range displacements the difference is even more evident. During this part of

the metering curve, the flow inertia is being opposing to the spool motion and the total

effect is a larger fluid dynamics loss. Similar behaviour can be seen also when considering

the discharge coefficients, see Figs. 25 and 26. When the flow rate is held constant, the

values during the first part of the opening travel are slightly larger than the ones during

Multidimensional Design of Hydraulic Components and Systems

269

Numerical

Inlet condition

Outlet condition

Spool positions

approach

Constant flow rate Constant pressure 0.038; 0.35; 0.68; Steady state

Case #1

(Q=12 l/min)

(pout=50 bar)

1.00

simulations

Constant flow rate Constant pressure 0.038; 0.35; 0.68; Steady state

Case #2

(Q=30 l/min)

(pout=50 bar)

1.00

simulations

Constant flow rate Constant pressure 0.038; 0.35; 0.68; Steady state

Case #3

(Q=100 l/min)

(pout=50 bar)

1.00

simulations

Constant pressure Constant pressure 0.038; 0.35; 0.68; Steady state

Case #4

(pin=70 bar)

(pout=50 bar)

1.00

simulations

Constant flow rate Constant pressure Function of time

Transient

Case #5

(Q=12 l/min)

(pout=50 bar)

(see Fig. 25)

simulation

Constant flow rate Constant pressure Function of time

Transient

Case #6

(Q=30 l/min)

(pout=50 bar)

(see Fig. 25)

simulation

Constant flow rate Constant pressure Function of time

Transient

Case #7

(Q=100 l/min)

(pout=50 bar)

(see Fig. 25)

simulation

Constant pressure Constant pressure Function of time

Transient

Case #8

(pin=70 bar)

(pout=50 bar)

(see Fig. 25)

simulation

Table 5. Boundary conditions used for the different simulated cases

closure, while as the direction reverses the discharge coefficients are larger due to the

inertia, but decreases remarkably as the metering edge reduces as an effect of the larger

losses. When comparing the discharge coefficients calculated by the steady state and

transient approach, the values are quite close, particularly at large spool displacement and for the opening travel. Conversely, if the values during the closure travel are considered, the transient simulations predict lower discharge coefficients as a results of the behaviour

mentioned before. Figs. 25 and 26 plot also the efflux angle for the different cases. The efflux angle is calculated accordingly to the reference system mentioned in section 4. A common

trend of the efflux angle is predicted for all the transient simulations. The angle tends to increase rapidly during the small spool displacements reaching a local maximum between

0.5 and 1.5 mm, ranging from small to large flow rate, while it decreases during the mid

range displacements and finally it starts increasing again at high spool axial positions. The efflux angle results to be slightly larger for the closure travel all through the displacements; the flow exiting the metering section during the closure phase results to be more disperse

when compared to the flow during the opening travel. This behaviour is clearly visible

when considering the flow field during the opening and closure travels, see Fig. 27. The

efflux angle trend described above for the transient simulations is definitively different from the behaviour predicted by the steady state calculations; in fact, with the steady state

approach the efflux angle is increasing as the spool displacements increases. Furthermore,

the efflux angle values calculated by this approach resulted to be larger than the ones obtained with the transient simulations. This difference has a remarkable importance if

the flow forces are accounted for, since they are significantly affected by the vena

contracta angle.

index-280_1.jpg

index-280_2.jpg

index-280_3.jpg

270

Applied Computational Fluid Dynamics

Fig. 24. Comparison between the a) pressure drop and b) flow rate through the metering

area predicted by the steady state and transient simulations.

Fig. 25. a) Discharge coefficients and b) efflux angle for the steady state simulations

Fig. 26. a), b) and c) Discharge coefficient and d), e) and f) efflux angle for the transient cases

index-281_1.jpg

Multidimensional Design of Hydraulic Components and Systems

271

Fig. 27. Velocity distribution for different non-dim. spool displacements (Case # 7)

6. Modelling approach for a multi-fuel injection system

The numerical approach described in the previous section for the design and performance

prediction of hydraulic components is employed for the analysis of a low pressure, common

rail, multi-fuel injection system. In particular, system is working with a mean pressure close to 3.5 bar and every injector is driven by a modified PWM control characterized by 3rd

order function opening and closing ramps and characteristic times adapted to the engine

rotational speed through an ECU correction. In the numerical analysis of this hydraulic

system, the turbulence as well as the cavitation models are used to address the flow within

the injector in order to estimate the permeability characteristics of the injector when

operating with different fuels. In the analysis, the operating conditions for the injectors are preliminarily evaluated by lumped and distributed parameter approach. Thus, the system

behaviour and the injection profiles for several fuel blends are calculated. The general

reliability of the one-dimension approach is defined with respect to experimental data

mainly in terms of injected mass per stroke. Fig. 28 shows the injection system model

including only the rail and the injectors. The connection to the supply is simplified by means of a flow-rate source. The Figure shows also the injector lift and the modified PWM wave;

the total time, tinj, is adjusted accordingly to the measurements. The CFD analysis focuses only on the injector geometry since it represents the main component influencing the

permeability of the hydraulic system. The injector is characterized by six nozzles: the first one is in-line with the injector axis, while the other five are located circularly every 72

degrees. Therefore, the injector geometry and the boundary conditions are cyclic

symmetrical allowing the simulations to be carried out only over a 72 degree sector mesh

(see Fig. 29). In the simulations the pressure at the nozzles’ exit is kept constant equal to 1

bar, while the inlet pressure is varied from 3.5, 5 and 10 bar. In fact, the lowest inlet pressure is the current injection pressure for this type of multi – fuel injection system, but the

tendency is to increase it in order to comply with the mass flow rate required by new fuels

and new injection strategies. Finally three different fuels are accounted for in the numerical analysis. Simulations with varying needle lifts and injection pressures are carried out for

pure ethanol, pure gasoline and a fuel mixture corresponding to 50% gasoline and 50%

index-282_1.png

index-282_2.jpg