Sie sind auf Seite 1von 76

2-way FSI simulations on a shock absorber check valve

M.Sc. Degree Project

Tommy Nilsson
tomnilss@kth.se
KTH Royal Institute of Technology
Department of Aeronautical and Vehicle Engineering
Stockholm, Sweden

Supervisors:
Matteo Pelosi
Ohlins Racing AB
Upplands Vasby, Sweden
Georg Ahrberg
Ohlins Racing AB
Upplands Vasby, Sweden

Examiner:
Stefan Wallin
KTH Royal Institute of Technology
Department of Mechanics
Stockholm, Sweden

January 7, 2015
Stockholm, Sweden
Abstract

A component of a hydraulic shock absorber, a check valve, is analyzed using


both numerical simulations as well as experimental testing. A fluid-structure
interaction (FSI) model is set up in ANSYS Workbench and is validated through
physical experiments - both steady state and transient. The fluid field is solved
in ANSYS Fluent and the structural deformation is solved for in ANSYS Struc-
tural. The coupling is made using ANSYS System Coupling.
The report covers the fundamentals of FSI analysis - methods of coupling fluid
and structure solution fields and methods for adapting the fluid mesh to account
for a changing geometry. A brief background on general moving/deforming
mesh algorithms are presented but the emphasis lies on the methods available
in ANSYS Fluent and how to apply these on the case of a shock absorber check
valve.
A moving/deforming mesh consisting of tetrahedral cells without inflation layers
on wall boundaries proves the most robust dynamic mesh setup. The exclusion
of inflation layers is shown to significantly affect the solution at low valve lift
height. At full lift the exclusion of inflation layers has no influence on the solu-
tion. The check valve is 4-fold axisymmetric but is shown to exhibit asymmet-
rical displacement. This is due to an asymmetrical fluid pressure distribution
on the check valve.
Steady state FSI simulations show satisfactory correlation to flow bench exper-
iments at low flow rates. The opening pressure differential of the check valve,
determined by the spring preload, is accurately predicted by the FSI model. At
flow rates above 10 l/min the differential pressure is under predicted, due to
simplifications to the computational domain.
Transient simulations and experiments both show an oscillatory pressure differ-
ential across the check valve as it opens, albeit with different frequencies.
Sammanfattning

En backventil tillhorande en hydraulisk stotdampare analyseras bade numeriskt


saval som experimentellt. En FSI-modell (Fluid-Structure Interaction) stalls
upp i ANSYS Workbench och valideras med fysiska experiment - bade statis-
ka och transienta experiment. Stromningsfaltet beraknas i ANSYS Fluent och
strukturdeformationer beraknas i ANSYS Structural. De bada losningar kopplas
sedan samman i ANSYS System Coupling.
Rapporten behandlar grundlaggande FSI - kopplingsmetoder och metoder for
rorliga berakningsnat. Generella metoder beskrivs kortfattat men fokus ligger
pa de metoder som finns att tillga i ANSYS Fluent, och hur dessa kan tillampas
i analysen av backventilen.
Ett rorligt berakningsnat bestaende av tetraeder utan inflationlager pa vaggrander
visar sig vara det mest robusta dynamiska natet. Uteslutandet av inflationla-
ger har dock betydande paverkan pa losningen vid laga ventillyfthojder. Vid
fullt oppen backventil har avsaknaden av inflationslager forsumbar paverkan
pa losningen. Backventilen ar 4-faldigt rotationssymmetrisk men oppnar anda
asymmetriskt. Detta forklaras av en asymmetrisk tryckdistribution pa backven-
tilen.
Statiska FSI-simuleringar visar pa tillfredstallande overenstammelse med
flodesbanksmatningar vid laga volymsfloden. Backventilens oppningstryck pre-
dikteras val av FSI-modellen. For volymsfloden over 10 l/min underbestams
tryckfallet pa grund av forenklingar av CFD-domanen.
Bade transienta simuleringar och transienta experiment visar pa oscillationer i
tryckdifferentialen over backventilen, om an med olika frekvens.
Preface

This thesis lying in front of you is the final proof of competence for obtaining a
M.Sc. degree in Aerospace Engineering from the Royal Institute of Technology
(KTH), Stockholm. The project has been carried out at Ohlins Racing AB in
Upplands Vasby, Sweden, under the supervision of Dr. Matteo Pelosi and Georg
Ahrberg. Examiner is Dr. Stefan Wallin from the Department of Mechanics at
KTH.
I want to thank my examiner Dr. Stefan Wallin for much appreciated feedback
and advice and I would like to thank my supervisors Dr. Matteo Pelosi and
Georg Ahrberg at Ohlins Racing for their support and guidance. I would also
like to thank Tobias Berg, ANSYS Sweden, for valuable technical support and
advice throughout the project.
Contents

1 Introduction 1
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.1 Ohlins TTX . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.2 Check valve . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Fluid-structure interaction . . . . . . . . . . . . . . . . . . . . . . 6
1.3 Project description . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.4 Similar studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2 Theory 11
2.1 CFD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2 FEM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.3 FSI . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.3.1 Energy conservation . . . . . . . . . . . . . . . . . . . . . 14
2.4 Dynamic mesh . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.4.1 Methods available in Fluent . . . . . . . . . . . . . . . . . 15

3 Method 17
3.1 Flow bench measurements . . . . . . . . . . . . . . . . . . . . . . 17
3.2 Damper dynamometer tests . . . . . . . . . . . . . . . . . . . . . 20
3.3 Fluent setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.3.1 Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.3.2 Mesh . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.3.3 Discretization . . . . . . . . . . . . . . . . . . . . . . . . . 29
3.3.4 Turbulence model . . . . . . . . . . . . . . . . . . . . . . 29
3.3.5 Mesh study . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3.3.6 Boundary conditions . . . . . . . . . . . . . . . . . . . . . 38
3.3.7 Fluid properties . . . . . . . . . . . . . . . . . . . . . . . 39
3.4 Structural setup . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.5 System Coupling setup . . . . . . . . . . . . . . . . . . . . . . . . 40
3.5.1 Steady state . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.5.2 Transient . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.6 Final transient model . . . . . . . . . . . . . . . . . . . . . . . . 43

4 Results 45
4.1 Flow bench measurements . . . . . . . . . . . . . . . . . . . . . . 45
4.2 Steady state FSI . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.3 Dynamometer tests . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.4 Transient FSI . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

5 Conclusion and Discussion 53


5.1 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
5.2 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5.2.1 2-way FSI as a development tool . . . . . . . . . . . . . . 56
5.2.2 Suggestions for further studies . . . . . . . . . . . . . . . 56

References 59

Appendix A UDFs and Profiles 63

Appendix B Model settings (360 case) 65


CHAPTER 1
Introduction

With computing power increasing exponentially with time according to Moores


law [1] the benefits of simulations as opposed to conducting physical experiment
are increasing as well. However, an inaccurate model will not produce more
accurate results by the use of more computing power. Here lies the scope of this
thesis - to set up and also validate a numerical simulation.

1.1 Background

Ohlins Racing has been developing high performance shock absorbers for the
vehicle industry since 1976. The development teams rely both on simulations
and experimental testing in the development of new products. A hydraulic
shock absorber works by the principle of controlling the oil flow rate through a
set of flow restricting valves. In Ohlins TTX models the damping characteristic
are determined by external valving and the flow path of the oil is controlled
by check valves. These check valves have been shown to also influence the
damping characteristics. Figure 1.1 shows the damper force measured in a
dynamometer in an experiment done prior to the start of this degree project.
The only difference is check valve setup, yet the damping characteristics are
different. This has inspired this FSI analysis of the check valve.

1.1.1 Ohlins TTX

A shock absorber of the conventional hydraulic type operates according to the


principle of converting kinetic energy to heat through viscous dissipation of
energy in a fluid. This is done by forcing a fluid through flow restricting valves,
which results in shearing of the fluid. The internal work done by the fluid
field working against its own internal shear stresses produces heat which is then
transferred, via the cylinder and piston, to the surrounding air.
Typically a shock absorber consist of a piston on a piston rod running inside a
cylinder, sealed on the one end and sealed around the piston rod on the other
2 CHAPTER 1. INTRODUCTION

400
3 N/mm, preload 3.9 N
8 N/mm, preload 10.4 N
200
Force [N]

200
0 0.5 1 1.5 2
Time [s] 2
10

Figure 1.1: Comparison in measured damper force between two


dampers with different check valve setups.

end. This is called a single tube damper. The cylinder is filled mostly with
viscous damper fluid but also a portion of nitrogen gas in a chamber separated
from the oil by a floating piston. As the damper is compressed the piston rod
will displace more of the internal volume in the cylinder. With the damper
being a closed hydraulic system, the internal volume of the system will then
decrease. Vice versa in rebound stroke, the internal volume of the damper is
increasing. This change in internal volume is taken up by the nitrogen gas, as
it is of much higher compressibility than the damper fluid.
The fluid is driven through port holes in the piston whereupon stacks of elastic
shims are placed (see Figure 1.2). These shims determine the opening pressure
and the opening characteristics of the flow ports. Flow in both directions go
through the piston. In compression stroke the rebound side shim stack is fully
closed and in rebound stroke the compression side shim stack is closed. The
pressure differential across the piston is responsible for the damping force.
Refering to Figure 1.3, a TTX damper hydraulic scheme is shown. A twin-
tube Ohlins TTX design presents many differences to a traditional single tube
damper. As mentioned earlier, on a conventional single tube damper the flow
ports and shim stacks are mounted on the piston. The pressure difference across
the piston drives the fluid through the valves in the piston. The twin tube TTX
damper works by to the same principle only the piston is solid and the valves
are external. Two sets of valves control compression and rebound damping
respectively.
In compression stroke the fluid will flow through the compression side valve
set and bypass the rebound valve set via a check valve. In rebound stroke
the flow goes the other way i.e. through the rebound valve set and around
the compression valve set via another check valve. The advantage to this design
with the solid piston and external valves is that the gas reservoir is connected to
the low pressure side of the fluid at all times. Comparing this to a conventional
damper where the gas reservoir is mounted on one side of the piston (typically
on the compression side [2, p. 4] which is the side that has the higher pressure
in compression stroke). This means that in compression stroke the flow through
the piston is driven by the low pressure side, since on the compression side the
1.1. BACKGROUND 3

External gas
Cylinder reservoir

Piston
rod

Figure 1.2: A conventional single tube damper. Source: Ohlins Racing.

gas will compress and thus no there will be no significant pressure build-up in
the fluid.
The problem here is cavitation. A fluid that is subjected to large pressure
gradients can, if the pressure is low enough, cavitate i.e. release gas bubbles of
air entrained in the fluid and further evaporate to gas state. When subjected
to higher pressure again, these gas bubbles implode generating shock waves,
which can lead to damages on nearby surfaces. To reduce the risk of cavitation
what is done is the gas pressure is set high enough to have the pressure on the
downstream side high enough to resist cavitation. On the TTX damper with
its external valve the gas reservoir is connected to the downstream side both in
compression and rebound stroke. The flow is then driven by the high pressure
on the upstream side and the downstream pressure is always equal to the set
gas pressure. With this design the gas pressure can be set lower than for the
conventional single tube damper. The lower gas pressure also enables the use
of looser fitting seals making for lower friction [2, p. 6].

1.1.2 Check valve

As stated above the damping in either direction is handled by two separate sets
of valves in a TTX damper. One for compression and one for rebound. The
valve set not in use is bypassed using a check valve. This is the hardware of
interest in this M.Sc. degree project. Its function is to fully restrict flow in one
direction and open to let the fluid pass freely in the other direction. Of course,
this component could also be designed to add additional damping, although its
4 CHAPTER 1. INTRODUCTION

Check
valve
open

Check
valve
closed

Check
valve
open

Figure 1.3: Flow schematics of a TTX damper of through-rod type.


Flow path in compression and rebound stroke. Source: Ohlins Racing.

main purpose is to control the flow path. The check valve is shown in Figure
1.5.

(a) Seen from the upstream (b) Seen from the downstream
side. side.

Figure 1.4: Check valve.

The pressure difference across the check valve forces it to open and the fluid
is allowed through. A flat shim, preloaded by a coil spring, rests on a seat
sealing off four port holes. The seat is slotted to reduce contact area. In the
check valve open state the fluid passes through the ports and through the gap
formed between the shim and the seat in an outwards radial direction. Figure
1.4 illustrates the design and function.
1.1. BACKGROUND 5

Flow path in main


flow direction

Shim
Coil spring
Spring collar
Lock ring
Flow path in reverse
flow direction

Figure 1.5: Check valve design and function (section view).

Throughout this report the fluid is referred to as upstream and downstream


of the check valve. Figure 1.6 shows the check valve mounted in a flow bench
fixture with the upstream and downstream sides marked.

Inlet
Upstream

Downstream

Outlet

Figure 1.6: Check valve mounted in the flow bench fixture.

The check valve has a very low opening pressure compared to the damping
valves. Its function is not to control flow rate and thus affect damping charac-
teristics but solely to set the flow path of the fluid. However, as the check valve
is not completely non-restricting to the flow some contribution to the damp-
ing will come from this component. Transient experiments done in the past
at Ohlins Racing have shown oscillations in the pressure difference across the
check valve. Figure 1.7 shows the oscillations in pressure across the check valve
on the compression side i.e. the check valve that is open when the damper is in
rebound stroke.
The shim is made out of the hardened and tempered carbon steel Sandvik 20C
[3]. It is characterized by good flatness, surface finish and high fatigue strength
under bending and impact stress. The same material is used for the shim stacks
where a lot of bending stress can occur. The shim in the check valve is unlikely
to see a lot of bending stress as the low preload of the coil spring allows for
translation of the shim rather than bending.
6 CHAPTER 1. INTRODUCTION

0.15
Oscillations in
pressure differential
0.10 over check valve.
p [MPa]
Compression
0.05 stroke Rebound
check valve stroke
closed
0.00
0 0.5 1 1.5 2
Time [s] 2
10

Figure 1.7: Dyno chart showing pressure drop over the check valve
in the compression side cavity. Spring rate: 3 N/mm, preload 3.9 N,
maximum lift height: 2 mm. Piston velocity is 50 Hz sinusoidal with a
maximum amplitude of 0.1 m/s.

1.2 Fluid-structure interaction

The analysis of a system consisting of a fluid and a solid can often be simplified
by making the assumption that the structural deformation is small and neg-
ligible. In cases where this assumption is no longer valid, a coupled analysis
might be imperative. Such is the case if the physical coupling between fluid
and structure is strong e.g. in the analysis of a sail on a sailboat. Here the
sail is the structure and the surrounding air is the fluid. The wind gives rise
to aerodynamic forces on the sail and as it is not rigid, it deforms. As the
sail is changing its shape the aerodynamic forces change as well, up until an
equilibrium is reached (given that the wind is constant and steady). This is an
example where the physical coupling of fluid and solid is strong. A change in
fluid flow gives rise to significant deformation of the structure and a change in
the geometry of the structure greatly affects the fluid flow.
In cases where deformations are small, the fluid-structure coupling is typically
weak e.g. a rotating boat propeller immersed in water. The propeller will induce
a jet of water behind it. The fluid is thus greatly affected by the structure. Not
by its deformation but by its rigid body motion. The structure is affected by
the forces induced by the fluid. There is a deformation but it is small and the
effect on the flow field of this small deformation is negligible. This is an example
of physically weak coupled interaction.
The interaction can be stable where the structural deformation will converge
towards a steady state deformation or it can be unstable. A good example of an
unstable interaction is wing divergence. A poorly designed airplane wing will
when deformed by the aerodynamic load increase its angle of attack and thus
give rise to increased load and in the worst case scenario break of completely.
The World War I fighter aircraft Fokker-D8 suffered numerous accidents due
to this phenomenon [4, p. 2]. The interaction can also by dynamic in nature
- dynamically stable or dynamically unstable. An interaction is dynamically
unstable when the oscillations of the structure is increasing in amplitude. Think
1.3. PROJECT DESCRIPTION 7

of an airplane wing again. When standing still on the runway a disturbance to


the wing structure will cause a vibration that quickly settles due to the internal
damping of the structure and to some extent the damping coming from the
surrounding air. When at speed, an induced disturbance will still be damped by
the structure but combined with unsteady aerodynamics the aerodynamic force
from the surrounding air might now add to the amplitude of the disturbance.
Even if the structure is statically stable it can still become dynamically unstable
and the oscillations will grow uncontrollably [4, p. 4].
There are several ways of modeling fluid-structure interactions, depending on
the physical nature of the problem. In the case of a strong physical coupling
the numerical coupling must also be strong. If the physical coupling is weak it
might be sufficient to solve the problem numerically using a weak numerically
coupled approach which would be less computationally expensive.
The modeling of the fluid and the structure can be done in several ways, and
depends on the geometry of the structure and the complexity of the the fluid
flow. Some cases can be successfully modelled using simplified linear models.
The case of a long slender airplane wing is such an example, where the fluid
forces can be modeled using potential flow theory and the wing structure can be
modelled using beam theory [4]. In the case of this M.Sc. degree project where
an opening and closing check valve is studied, the fluid flow is solved for using
finite volume CFD and the structural deformation will be solved for using the
finite element method. The coupling of the two solution fields is done using a
partitioned iterative 2-way method.

1.3 Project description

The goal of this M.Sc. degree project is to set up a robust and accurate 2-way
FSI model in ANSYS Workbench for simulating the transient behaviour of a
check valve. The solution is to be validated with flow bench measurements and
shock absorber dynamometer tests.
A simplified domain is used for the majority of the work as this allows simula-
tions to be run on a laptop computer. A final larger model, based on lessons
learned from the simplified model, is in the end stage of the project, run on a
computer cluster.

1.4 Similar studies

In 2003 Leon Leventhal published a study [5] on a ball check valve using both
experimental methods and FSI. In the study the author is able to resolve the
dynamic opening phase of the check valve as well as bouncing effects when
closing. The FSI tool used is ADINA [6] which is a direct-coupled FSI code.
The fluid and structural models are 2D axisymmetric. A small gap of 20 m
is left between ball and seat in its closed state to not pinch off the fluid mesh.
In the structural model, the coil spring is modelled using a spring element.
The paper does not include much detail about the CFD model or what method
8 CHAPTER 1. INTRODUCTION

is used to deform the mesh, and although experiment are conducted there is
no direct comparison to the FSI model for validation. Figure 1.8 shows the
axisymmetric 2D fluid domain of the study.

Figure 1.8: CFD domain in Leventals study [5].

Parameter studies are done both experimentally and using the FSI model and
both studies lead to the same conclusion. Lower spring preload reduces the
initial pressure spike and oscillations in the opening phase but can cause ball
bounce at closing. FSI studies showed that spring rate had minimal influence
while experiments gave inconclusive results. The check valve in this study was
of ball-in-seat type. Reducing seat area showed lower pressure spikes and less
oscillations at check valve opening. This effect is explained as an effect of a
smaller throttling area between ball and seat and thus less influence by Bernoulli
effects. It was also shown that a ball with a diameter as small as the seat inner
diameter made for smaller pressure spikes.
Price et. al of Engineering Dynamics Inc, Texas [7] showed in a study of a
check valve used in a reciprocating pump the effects of modifying shim and
seat geometry. The check valve is similar to the one studied in this M.Sc.
degree project, a flat shim resting on a seat and preloaded by coil springs,
albeit pressures and length scales are about one order of magnitude larger in
the study by Price et. al. Figure 1.9 shows the check valve of their study.
They refer to the adhesion force between shim and seat in the opening phase
as stiction (short for static friction) and in their study they found the stiction
force to be a function of the width of the seating surface, the viscosity of the
fluid and the initial film thickness which depends on the surface finish. They
model the stiction phenomenon by considering two flat plates immersed in a
fluid and as the two surfaces are separated fluid fills the void and in doing so
generates a force on the surfaces in accordance with Bernoullis principle. The
static pressure in the high velocity fluid filling the void is lower than in the
surrounding fluid. In the case of their check valve this stiction phenomenon was
enough to cause cavitation leading to damage on the shim and seat surface. To
reduce seat area they machined grooves into the seat, which effectively reduced
the stiction force and also increased the flow rate of the pump the check valve
1.4. SIMILAR STUDIES 9

was used in. This was due to reduced valve lag in the pump. It was also
shown that a fluid with lower viscosity gave rise to smaller pressure spikes.

Figure 1.9: Check valve used in the study by Price et. al [7].
CHAPTER 2
Theory

The term FSI encompasses a number of different methods. The solution can
be steady-state or transient, the coupling can be weak or strong, the solid so-
lution can be structural deformation, temperature or both. The fluid field can
be solved for using simple methods like potential flow, or computationally ex-
pensive methods such as DNS. In this M.Sc. degree project the solid solution
is structural deformation solved for using the finite element method in ANSYS
Structural - both static and transient. The fluid field is solved for using finite
volume CFD in ANSYS Fluent. The two solution fields are coupled using a
partitioned iterative method in ANSYS System Coupling. Below follows a brief
theoretical background of the solution methods, and coupling method used.

2.1 CFD

Computational fluid dynamics (CFD) is the name for a number of methods when
it comes to solving a fluid flow field numerically. In common for most CFD
method are the Navier-Stokes equations1 . This set of equations can describe
any fluid flow. There is yet no mathematical proof that a solution exists to the
Navier-Stokes equations for arbitrary boundary conditions. This is part of the
Millennium problem2 . In practice however, with proof or not, the equations
can be solved. The equations govern the conservation of mass, momentum and
energy. In the case of finite volume CFD, the equations are expressed as volume
integrals and solved on a finite volume mesh. The mesh resolution governs
the length scale of flow features that can be resolved. On turbulent flows,
solving Navier-Stokes equations on a mesh fine enough to resolve the smallest
length scales, the chaotic fluctuations that is turbulence, is computationally very
expensive and is yet only relevant for academic purposes. This would be, what
is known as DNS - Direct Numerical Simulation. Instead what is commonly
1 There are also the Euler equations for inviscid flow and lattice Boltzmann equation used
in codes such as PowerFlow [8] and XFlow [9].
2 One of the Millennium problems is to derive proof of existence and smoothness of

NavierStokes solutions for arbitrary boundary conditions. The prize for doing so is $1,000,000
[10, p. 57].
12 CHAPTER 2. THEORY

done is to resolve mean velocities and model fluctuations, as done in RANS


(Reynolds-Averaged Navier-Stokes) CFD.
In between DNS and RANS on a scale of computational cost lies LES (Large
Eddy Simulation). In LES the smallest turbulent length scales (sub-grid scales)
of the flow field are modelled while the larger length scales are resolved on the
grid. In RANS CFD, also the larger length scales are modelled.
The idea behind RANS is to separate the flow velocity into an average and a
fluctuating part. The fluctuating part, the Reynolds stress needs to be modelled
to close the system of equations. The models for doing so are called turbulence
models, and involve a statistical averaging procedure of the Reynolds stress [11].
The use of a turbulence model to model the smallest scales of the flow field and
resolving the larger scales can be an effective way to solve a flow field. It is
however important to use a turbulence model suited for the particular flow that
is studied, as no turbulence model is universal and suited for each and every
flow problem.

2.2 FEM

The finite element method (FEM) is a numerical method of approximating the


solution to a differential equation on a finite element grid. The method was
initially developed for stress and displacement calculations for structural anal-
ysis [12, p. 429] but is not limited to solid mechanics. The same discretization
method can also be used for solving thermal, fluid and electrostatic problems.
In the scope of this thesis the use is limited to solid mechanics. The domain
is divided into finite elements connected to each other in nodes. Each element
is represented by an algebraic equation relating stiffness and displacement to
force making for a system of equations to be solved for the whole structure.
Depending on the type of element used different problems can be solved. Com-
pare this to the finite volume discretization of finite volume CFD, where each
finite volume acts as a control volume but does not include any details about
the physics of the fluid. In finite element analysis it is the choice of element
type that govern the physics of the problem. For structural problems there are
primitive elements such as bar and beam elements applicable to simpler prob-
lems as well as continuum elements such as plates and solids for discretizing
more complex problems.

2.3 FSI

As described earlier in the introduction, the physical coupling between solid and
fluid can be weak or strong. So can the numerical coupling between the flow
solution and the structural solution. 1-way FSI is a weak numerical coupling.
This is where data is transferred in only one direction. For example the flow
solution is calculated and sent to the structural solver which solves for a defor-
mation, whereupon the flow solver calculates the solution for the next time step
and the process continues. This is applicable only to cases that are physically
2.3. FSI 13

weakly coupled, as in the example of boat propeller used earlier. Figure 2.1a
illustrates the principle.
There are two types of strong coupling methods - monolithic and partitioned
iterative. Monolithic FSI, or direct-coupled FSI, means that both deformation
and flow is solved directly using the same solver. Software such as ADINA
[6] and COMSOL Multiphysics [13] have these capabilities. The monolithic
approach requires more storage but is better suited than the partitioned ap-
proach to problems with large nonlinearities [14, p. 1442]. For large problems,
however, the partitioned approach is often the preferred method [15]. Solv-
ing complex structural problem using the monolithic approach is in general
computationally challenging, mathematically and economically suboptimal and
software-wise unmanageable [15].
The partitioned iterative approach uses separate solvers - one for the flow field
and one for the structural deformation. The solution is transferred in both
directions. Figure 2.1b illustrates the principle. During each time step solution
data is transferred back and forth between the solver in coupling iterations until
the coupling step is converged, that is when the data transfer is low enough
to meet convergence criteria. This is 2-way FSI. The most popular solution
procedure is referred to in [15] as the Conventional Serial Staggered procedure
and goes as follows:

1. Transfer the motion of the FSI boundary in the structural domain to the
fluid domain,
2. Update moving mesh in the fluid domain,
3. Advance in time and compute fluid solution,
4. Convert pressure and fluid stress into structural loads,
5. Advance in time and compute structural solution.

time step n time step n + 1 time step n time step n + 1

Fn F n+1 Fn F n+1

Sn S n+1 Sn S n+1

(a) 1-way coupling. (b) 2-way coupling.

Figure 2.1: Coupling types. F is fluid solution and S is solid solution.

When making multiple coupling iterations per coupling step it is not necessary
to reach full convergence for the fluid solution in every coupling iteration, as
long as the final coupling iteration is converged. It can actually be a good idea
to not converge the fluid solution fully in each coupling iteration as this will
provide some damping to the solution.
14 CHAPTER 2. THEORY

For weakly coupled systems one stagger loop (coupling iteration) per time step
might be enough. This would be explicit iterative FSI, and would not be suited
for cases of strong physical coupling, unless it is only the steady state solutions
that is of interest. An explicit iterative coupling scheme does not achieve a full
fluid-structure coupling in each time step and is in general energy increasing and
hence numerically unstable [16, p. 2-3]. Of the monolithic and the partitioned
approach, monolithic FSI is the strongest numerical coupling. Note that if
each coupling step in the partitioned solution procedure is fully converged a full
coupling between fluid and solid is achieved, and the solution should in theory
(if the data transfer residuals are infinitesimal) be the same.
For this M.Sc. degree project the partitioned iterative approach is used, in
the form of ANSYS Fluent as the fluid solver and ANSYS Structural as the
structural solver. ANSYS System Coupling handles the data transfer between
the two solvers.

2.3.1 Energy conservation

An FSI system consist of two subsystems - a fluid system and a structural sys-
tem. The boundary condition that couples fluid system to structural system is
the FSI interface. On this interface pressure loads from the fluid are transferred
to the structural domain and causing a deformation. Vice versa, deformations
are passed back to the fluid system giving rise to forces on the fluid. As the two
subsystems together form a closed system, they share a global energy budget.
As the FSI interface moves the solid performs work against the internal forces
of the fluid, producing a change in its energetic status [17, p. 342]. To study
transient effects in a fluid-structure interaction it is important that the coupling
scheme is energy conservative. Especially when dealing with systems of unstable
nature, such as in, for example, the study of wing flutter. When using FSI to
find limits of stability for such problems it is important to use a method that
is as globally conservative as possible to not influence the solution by adding or
taking away artificial numerical damping [15].
Michler et. al [16] compares, in a numerical experiment, solution accuracy and
stability using conservative and non-conservative monolithic schemes. They
show how a scheme that conserves mass, momentum and energy at the fluid-
solid interface results in a more accurate solution at the same computational
expense, compared to a non-conservative scheme. They also show that violating
energy conservation at the boundary can lead to numerical instabilities, even
for a monolithic scheme.
In ANSYS System Coupling the force mapping between the two meshes on the
FSI interface is locally (on element level) and globally conservative [18, lecture
2, p. 13]. The two meshes, the fluid mesh and the solid mesh, does not need to
match on the boundary. Typically the fluid mesh is finer than the solid mesh
[19, p. 40], but it is recommended to have similar length scales on both meshes
to maintain load resolution.
System Coupling does not make sure that the global energy of the fluid-structure
is conserved - the scheme is non-conservative w.r.t. energy.
2.4. DYNAMIC MESH 15

2.4 Dynamic mesh

In FSI one important part of the problem setup is the dynamic mesh. As for
a fixed geometry CFD-only calculation, the quality of the mesh is crucial in
obtaining an accurate solution. In the case of FSI with transient boundary
conditions and moving boundaries, this becomes a sometimes difficult task. In
this particular case of a check valve the geometry goes from a very narrow gap
only about 20 m to a maximum gap of 1 mm in its open state. Meshing these
two geometries separately as fixed geometries is no difficulty. The difficulty
comes in setting up a dynamic mesh that can adapt to the moving geometry
while still maintaining good quality throughout the simulation.
There are two types of dynamic mesh methods. A moving mesh where the mesh
tracks and follow the moving boundary and deforms accordingly. The other type
is the non-moving (also known as overlapping or overset) mesh or interface-
capturing method. With the interface-capturing method the moving boundary
is intersected with the overlapping mesh and the resolution at the boundary is
determined by the overset mesh resolution and not the resolution of the actual
moving boundary. For all cases where a moving mesh is applicable the same is
also recommended rather than a non-moving mesh because of better accuracy
near the FSI interfaces [20, p. 84]. Some cases are suited for a combined use of
both methods. Such an example is water splashing where the bulk of the water
is tracked by a moving mesh while splashing droplets are captured using the
interface-capturing method [21].

2.4.1 Methods available in Fluent

In the case of the check valve, the geometry is simple enough to allow the use of
a moving mesh. Fluent has a few different ways of moving and deforming the
mesh. Available methods are Smoothing and Remeshing. Both methods belong
to the moving mesh category described above. Smoothing essentially relocates
the nodes to adapt to a moving boundary. The boundary nodes follow the mov-
ing interface while the interior nodes are moved according to a chosen algorithm
- Spring Smoothing or Diffusion Smoothing. Both Smoothing methods work in
a similar fashion. Diffusion Smoothing is a bit more computationally expensive
but does tend to generate a better quality mesh and is recommended in cases
with large deformation and is therefore the method of choice in this case [22].
There are benefits to using Spring Smoothing such as the possibility to have a
prism layer mesh that moves rigidly with the moving boundary. More on this
in following chapter.
Remeshing works, in Fluent, by a set of size and skewness constraints, when
smoothing the mesh stretches or compresses cells too much or cells reach a
skewness limits, local remeshing is activated. Here nodes are reconnected. Cells
are either split or merged. When adapting the existing mesh using Smoothing
and Remeshing is no longer possible while maintaining size and skewness con-
straints a total remeshing of the moving mesh is triggered. This is in Fluent
called Cell Zone Remeshing and creates a new mesh from scratch using the cell
size of the existing mesh. Usually when meshing a geometry such as this, local
16 CHAPTER 2. THEORY

refinement are made where velocity gradients are high and a coarser resolution
is used in non-critical areas. When using Remeshing, both local and the global
Cell Zone method, the meshing algorithms work best on a mesh of uniform el-
ement size. Local Remeshing works by set size limits and Cell Zone Remeshing
calculates the element size for the new mesh from the existing mesh. Thus hav-
ing local refinements will affect the average element size, and will thus make for
an increased number of elements after every Cell Zone Remeshing.
With Smoothing all types of elements can be used, while Remeshing works
only with tetrahedral elements. The global Remeshing algorithm Cell Zone
Remeshing has the capability of remeshing both tetrahedrons and prisms. This
is however a computationally expensive method compared to local Remeshing,
and should thus only be used when local Remeshing fails to maintain size and
skewness criteria.
There is also a third type of dynamic mesh method - Layering. This method is
commonly used when modelling in-cylinder flows in combustion engines. The
downside is that it can cause instabilities when used on strongly coupled FSI
cases such as this check valve case. When the FSI interface moves mesh layers
will be added or subtracted at the interface boundary, and this might cause error
to amplify and cause instabilities within the coupling step. For the purpose of
the check valve setup modelled in this thesis the Layering method is deemed
unfit because of this possible instability issue.
CHAPTER 3
Method

The end goal is to have a working FSI model for simulating transient check valve
behaviour. This model is set up and validated in steps. First a steady state
model is compared to quasi-steady state flow bench measurements and later
a transient model is set up and compared to transient damper dynamometer
tests. This chapter documents the setup and and execution of experiments and
model.

3.1 Flow bench measurements

The flow bench consist of a pump driving damper fluid through a set of pipes
leading up to a chamber where the specimen to be measured is placed. Figure
3.1 shows the flow bench and Figure 3.2 the check valve fixture. The check valve
is mounted in a fixture with the inlet being on the side and the outlet on the
bottom. Pressure sensors are mounted upstream and downstream of the fixture
as shown in Figure 3.1.

Pressure sensor,
downstream

Pressure sensor,
upstream

Figure 3.1: Picture of the flow bench used for steady state experiments.

The flow bench can be configured to run in both directions. In this configuration
18 CHAPTER 3. METHOD

Figure 3.2: Fixture for mounting the check valve in the flow bench.
Inlet is on the side and outlet at the bottom. The top is sealed after the
check valve has been positioned.

the upstream pressure is measured in the large chamber holding the fixture. In
Figure 3.3, a schematic overview of the flow bench is shown. Because of the large
size of the measurement chamber, the flow close to the upstream pressure sensor
is assumed stagnant and the measured pressure is seen as the total pressure
upstream of the check valve. The pressure sensor downstream of the check
valve is connected to the outlet piping downstream of the fixture outlet. The
pressure here can not be assumed as the total pressure.

pup
Check valve
Measurement
chamber pdown

Pump

Figure 3.3: Flow bench schematic.

According to Bernoullis principle total pressure is


1
p0 = p + U 2 + gz (3.1)
|{z}
total
|{z}
static
|2 {z } pressure
|{z}
pressure pressure dynamic level
pressure

where p is static pressure, is fluid density, U is velocity, g is acceleration due to


gravity and z is elevation. The last term can be omitted in this case. To relate
it in numbers, the maximum velocity (cross-sectional average) at the fixture
outlet is
Qmax
Vmax = = 4.9 m/s (3.2)
Aout
where Qmax = 7.5 104 m3 /s (45 l/min) is the maximum flow rate used in
model validations later, Aout = 1.5 104 m2 is the area of the fixture outlet
making for a dynamic pressure of
1 2
pdyn = V = 0.010 MPa (3.3)
2 max
3.1. FLOW BENCH MEASUREMENTS 19

where = 843 kg/m3 is fluid density at 25 C. The dynamic pressure is not


negligible considering the order of magnitude of the measured pressure drop at
the same flow rate is O(0.1 MPa).
The measured pressure drop is the difference in total pressure upstream and
static pressure downstream,
p = pin pout + pcalib.error = pin
0 p
out
(3.4)
since the flow upstream is assumed stagnant.
The pressure drop across the check valve is measured at different flow rates
to produce what is know as a p Q diagram - steady state pressure drop
against volumetric flow rate. Flow rate is increased linearly with time while the
pressure upstream and downstream of the check valve is measured. The flow
rate is increased at a rate slow enough to omit transient effects. By measuring
pressure drop both on flow rate ramp-up and ramp-down and comparing the
data quasi-steady state behaviour can be confirmed.
At zero flow rate there is a measured pressure drop, probably due to a calibration
error (p 6= 0 at Q = 0). The following measurements are therefore corrected
with this calibration offset shown in Figure 3.4.

102
1.0
Pressure drop [MPa]

0.5

0.0
Calibration
0.5 error

0 2 4 6 8 10 12
Flow rate [l/min]

Figure 3.4: pref vs. volumetric flow rate for one of the tested check
valve setups.

For the purpose of validating the CFD model alone a series of measurement are
made without the coil spring present, but with different maximum lift heights.
Without the coil spring the shim lift will be equal to its maximum permitted
lift for the full duration of the flow rate ramp. With the lift height known
fixed geometry CFD cases can be set up to validate the CFD model alone.
Two spring collars of different lift height, 1 mm and 2 mm, combined with thin
spacers allows the lift height to be changed in small increments. The volumetric
flow rate is ramped from zero up to 45 l/min over a period of 30 seconds. The
flow rate is then decreased at the same rate down to zero. With this rate of
change steady state behaviour is confirmed as the the p Q relation is the same
in ramp-up as ramp-down as shown for one of the check valve setups in Figure
3.5.
To validate the FSI model measurements are made with the coil spring installed.
mp-down
20 CHAPTER 3. METHOD

.05 0.15
Ramp-up

Pressure drop [MPa]


Ramp-down
0.10

0.05

0.00
0 10 20 30 40
Flow rate [l/min]

Figure 3.5: p vs. volumetric flow rate in ramp-up and ramp-down for
the case of a 3 N/mm spring and 3.9 N of preload.

Two different spring rates are tested in combination with the two spring collars

.00
giving a total of four setups (Table 3.1). There is also the possibility of changing
spring preload. This option is left out for now.

Table 3.1: Flow bench test setups.

0 20
Flow
Flow
Flow
bench
bench
bench
setup
setup
setup
1:
2:
3:
10
Spring rate [N/m]
3
8
3
Preload [N]
3.9
10.4
3.9
Max. lift [mm]
1
1
2
Flow bench setup 4: 8 10.4 2

3.2
Flow rateDamper dynamometer tests

20 30 40
In transient damper dynamometer tests on a motorcycle front end, carried out
before the start of this M.Sc. degree project, oscillations in the pressure dif-
ference across the check valve could be seen (Figure 1.7). If these pressure
oscillations are effects of the check valve or effects of other flow restrictions in
series is hard to say. In an attempt to isolate the effects of the check valve an
other test is conducted. This time with an empty check valve and no shim stack
in the rebound cavity, and pressure drop measured over the compression side

Flow rate [l/min]


check valve.
For this test a damper dedicated for testing, with outlets for pressure sensors,
is used (Figure 3.6). Driving the fluid is the piston and cylinder of a TTX36
shock absorber. Specifications for the experiment damper are listed in Table
3.2.
In rebound stroke the fluid passes through the empty valves on the rebound
side, over to the compression side cavity and through the check valve. Pressure
is measured upstream and downstream of the compression side check valve. The
sensors are mounted flush with the flow boundaries and the cross-sectional area
is the same at the three sensor locations. Figure 3.7 shows the location of the
3.2. DAMPER DYNAMOMETER TESTS 21

Figure 3.6: Dyno damper.

Table 3.2: Dyno damper dimensions.

Piston diameter [mm]: 36


Piston rod diameter [mm]: 12
Valve cavity inlet/outlet diameter [mm]: 14

pressure sensors. Only two of the pressures are of interest here - upstream and
downstream of the check valve when the shock absorber is in rebound stroke.
As both inlet and outlet to the check valve cavity are of the same cross-sectional
area there will likely not be any difference in dynamic pressure due to different
flow velocities. The measured difference in static pressure is therefore assumed
equal to the total pressure drop,

pin out in
dyn = pdyn p = p p
out
= pin out
0 p0 (3.5)

Empty damping
Damping valve
valve
Pressure sensor
Gas reservoir
upstream
Empty check
Check valve
valve
Pressure sensor Flow direction in
downstream rebound stroke

Figure 3.7: Dyno damper end piece (section view).

Three check valve setups are tested, listed in Table 3.3. Preload is changed by
placing shims between the lock ring and spring collar. In Setup 2 the spring is
preloaded with two of these shims (0.3 mm thick each). The maximum lift is set
by the spring collar. To not make the check valve bottom out in Setup 2 with
0.6 mm of added shims the lower spring collar (2 mm maximum lift) is used
for this setup making the maximum lift 1.4 mm. The difference in maximum
lift between Setup 1 and Setup 2 should have no influence as the check valve is
unlikely to open fully. Steady state flow bench tests indicate that the check valve
22 CHAPTER 3. METHOD

in its most compliant setup (3 N/mm spring rate and 3.9 N of preload) reaches
1 mm lift at about 30 l/min and the maximum flow rate in the dynamometer
tests that will be looked at is about 15 l/min. The check valve lift might exhibit
some overshoot in the transient tests and it is not impossible that the check
valve reaches maximum lift, but it is unlikely. If so, it should be visible in the
measured pressure drop vs. time diagram.

Table 3.3: Dynamometer test setups.

Spring rate [N/mm] Preload [N] Max. lift [mm]


Dyno setup 1: 3 3.9 1.0
Dyno setup 2: 3 5.9 1.4
Dyno setup 3: 8 10.4 1.0

The test case used for FSI model validation is a sinusoidal piston velocity with
frequency 50 Hz and maximum piston velocity 0.25 m/s. In earlier tests this is
where oscillations in pressure were most apparent. The dynamometer has sen-
sors for logging position. Figure 3.8 show the position and velocity (calculated
from position) for the test program used for FSI validation. Velocity is defined
positive in compression stroke and negative in rebound.

102 0.250
1.15
Velocity [m/s]

1.20 0.125
Position [m]

Re
.
mp

b.

1.25 0.000
Co

1.30 0.125

1.35 0.250
0 0.5 1 1.5 2 0 0.5 1 1.5 2
Time [s] 102 Time [s] 102

(a) Dyno position. (b) Dyno velocity.

Figure 3.8: Dynamometer states. Dyno position is measured while


velocity is derived from position and time.

3.3 Fluent setup

The end goal is to set up a robust and accurate transient FSI model. As FSI is in
general computationally expensive, compared to solving a standalone fluid case
or a standalone structural case, it is worth-wile to simplify the case if possible.
At least so during the setup and debug phase. Here a smaller simplified domain
is used for the larger part of the work when investigating mesh size, dynamic
mesh settings, turbulence modelling and time step size. In the final stage of the
project the size of the domain is increased to better represent the real geometry
used in experiments.
3.3. FLUENT SETUP 23

3.3.1 Geometry

For studying the impact of mesh size, turbulence model, time step etc. the do-
main is simplified to allow the case to be run on a laptop computer with reason-
able computation times (overnight). The geometry upstream and downstream
of the check valve is that of the fixture used for the flow bench measurements,
with the difference that the inlet is now on the same axis as the outlet making
the geometry 4-fold axisymmetric. The outlet is extended to avoid reversed
flow at the boundary, which can cause convergence issues. Figure 3.9 shows
the actual geometry of the flow bench fixture and the simplified geometry. To
reduce the size of the domain further only a 45 slice of the geometry is used as
computational domain.

Inlet

Outlet

(a) Flow bench (b) Simplified (c) 45 slice.


fixture geometry geometry with
(section view). extended outlet
and axisymmetric
inlet (section
view).

Figure 3.9: CFD geometry.

The CFD domain for the FSI case does not include the entire length of piping
in the flow bench leading up to the downstream sensor. The CFD domain is
also simplified upstream and does not include the large cavity where the up-
stream pressure sensor is mounted, but any pressure drop between the upstream
pressure sensor and the fixture inlet is assumed small and negligible.
The same geometry is used for the steady state simulations as for the first
transient simulations.

3.3.2 Mesh

The fluid domain is separated into smaller subdomains (Figure 3.10) to create
sweepable bodies where applicable, as the number of nodes and cells for a swept
body is typically much smaller than ones meshed with the default mesher [23, p.
231]. The regions connecting the swept regions are meshed with unstructured
tetrahedrons as a sort of buffer zones, connecting the swept mesh zones. The
region surrounding the moving shim is separated from the rest of the mesh.
24 CHAPTER 3. METHOD

This is to make the size of the dynamic mesh domain as small as possible for
computational reasons. The dynamic mesh operations takes place on a single
core and can be a bottleneck in the solve process when the case is scaled up
on multiple cores. Care is therefore taken to make the dynamic mesh region as
small as possible.

Swept hex.
Inlet
mesh

Deforming
mesh zone

Unstructured
Outlet
tet. mesh
(a) Simplified domain. (b) Mesh zones.

Figure 3.10: Simplified geometry and mesh.

Dynamic mesh

Maybe the most time consuming and difficult part of setting up this FSI model
is the dynamic mesh. First of all, the fluid mesh needs to be continuous thus
a small gap that can fit at least one cell needs to be left between the shim and
the seat. A 20 m gap is left. Similar work done on a ball check valve showed
that this was small enough to capture Bernoulli effects1 [5].
Ideally the mesh would include inflation layer along all wall boundaries to better
resolve the boundary layer. In the initial narrow gap these prism cells would
be very thin. As the check valve opens the prism cells would ideally stretch, or
additional layers of prism cells would be added. When it comes to cell size, the
mesh would ideally be finer in regions with large gradients in pressure and/or
velocity and around boundaries with small radius of curvature and could be
made coarser in less active flow regions. In the case of fixed geometry CFD this
is usually not a problem, but when dealing with a transient FSI case with large
deformations compromises in mesh quality might have to be made.
As mentioned in the Theory section, there are two methods available in Fluent,
applicable to this case, for adapting the mesh - Smoothing and Remeshing. The
goal is to have a dynamic mesh that is robust and can handle going from a very
narrow gap to fully opened.
When investigating different types of dynamic mesh setups a prescribed motion
is assigned to the shim using a Fluent Profile (Appendix A). The mesh deforma-
tion can then be previewed without solving for the flow field. This proved to be
a very useful feature during the setup. Both axial as well as some radial motion
1 At check valve opening the fast flowing fluid in the gap gives rise to a low static pressure

on the shim which delays check valve opening.


3.3. FLUENT SETUP 25

of the shim was tested. As a simplified problem the shim can be constrained
to axial motion only, but the ambition is to eventually replace the ideal spring
element in the structural model with a model of the actual coil spring which
will most likely lead to some radial motion as the reaction forces in a coil spring
are usually not perfectly axial.
Initially simple 2.5D2 cases where studied to get an idea of the influence of the
different Smoothing and Remeshing parameters, but this approach was quickly
abandoned for 3D cases since the dynamic meshing works differently on faces
and interior volumes. Many different approaches were tried and a brief summary
of each approach follows below.

Dynamic mesh approach #1 - Unstructured tetrahedral mesh with


inflation layers in the same domain:

This first mesh attempt consist of inflation layers along all boundaries and a
tetrahedral mesh in the rest of the deforming zone. The inflation layers were
made very thin in the gap. The problem is that the prism cells, which are a
lot smaller in size than the tetrahedrons is included in the cell size calculation
done by the Cell Zone Remeshing algorithm and thus for every time Cell Zone
Remeshing is triggered the cells will decrease in size and in the end the mesh will
be very dense. Figure 3.11 show the mesh after Cell Zone Remeshing has been
triggered. Initially the interior cells had the same dimensions as the face mesh
(in blue). Disabling Cell Zone Remeshing and relying only on local Remeshing
was tried but caused problems of negative cells.

Small cells

Figure 3.11: Mesh after Cell Zone Remeshing. Inflation cells and tet.
cells in the same cell zone make for an increase in number of cells every
time Cell Zone Remeshing is triggered.

An other issue is the splitting of the cells in the shim-to-seat gap. If the check
valve opens too fast i.e. the time step is too large, the cells in the gap tend
to stretch and become larger than the size constraints set. Figure 3.12 shows
stretched cells in the gap. The mesh shown in the figure is in 2.5D i.e. only
one element deep. On a 3D mesh the effect of stretched cells is not nearly as
prominent. This goes to show how the dynamic mesh works differently on face
cells and interior cells.
2A 2D mesh extruded one cell deep is usually referred to as 2.5D.
26 CHAPTER 3. METHOD

Stretched cells

Figure 3.12: Stretched cells in gap. Here on a 2.5D mesh.

In the early stages of exploring dynamic mesh settings, the stretched cells ap-
peared to be a problem also on interior cells of a 3D mesh. By manually trig-
gering Cell Zone Remeshing at predefined times as the check valve opens the
stretched cells were be forced to split. However, after running the FSI case and
monitoring mesh deformation this method of manually forcing the cells in the
gap to split proved unnecessary. The short time step used for the transient
simulation eliminated the problem.

Dynamic mesh approach #2 - Structured mapped mesh:

A mesh of hexagonal cells is mapped as shown in Figure 3.13a. Using a mapped


hexagonal mesh precludes the use of Remeshing. Instead the mesh adaptation
must rely solely on Smoothing. Good mesh quality is kept in the gap, but
the problem is skew cells around the shim edges. After trying a multitude
of Smoothing settings, Spring based and Diffusion based, without success this
approach was ruled out. With both Smoothing methods negative cells were
encountered at the shim edges as shown in Figure 3.13b.

(a) Undeformed mesh. (b) Negative cells.

Figure 3.13: Dynamic mesh approach #2. Mapped hexagonal mesh.

Dynamic mesh approach #3 - Mapped mesh in gap and unstructured


tetrahedrons in the rest of the domain:

To have good mesh resolution in the gap while still having a coarser and also
more robust dynamic mesh in the rest of the domain, the domain is separated
3.3. FLUENT SETUP 27

into four subdomains. The region between shim and seat is mapped with hexag-
onal cells while the rest of the domain is meshed with unstructured tetrahedrons.
Smoothing is used in the hexagon regions. Since no cells will be added as the
gap increases the number of hexagonal cells must be enough to provide satis-
factory resolution also in the fully open state. Smoothing in combination with
Remeshing is used in the tetrahedron regions. The nodes are unmatched on
the interface, as matching the cells would cause very small tetrahedral cells
(which works poorly with Remeshing). The deformation of the tetrahedral cells
is satisfactory. The uniform cell size of the tetrahedral mesh makes both local
Remeshing and Cell Zone Remeshing work to satisfaction. The reason for aban-
doning this approach is the restriction in degrees of freedom of the shim. The
dynamic mesh work well for axial-only translation, while radial translation pro-
duces skew hex cells at the mesh zone interface. Figure 3.14 shows the deformed
mesh.

(b) Closeup of skew cells at


(a) Deformed mesh. interface.

Figure 3.14: Dynamic mesh approach #3. Hexagonal mesh in the gap
and unstructured tet. cells in the rest of the dynamic mesh zone.

Dynamic mesh approach #4 - Unstructured tetrahedral mesh with


inflation layers in separate domains:

This mesh is similar to approach #1 but now with the inflation layers in their
own separate cell zones (Figure 3.15). This makes the Remeshing of the tetra-
hedrons work better. The problem with Cell Zone Remeshing encountered in
approach #1 is now gone. That is the problem with an underestimated cell size
for the tetrahedral cells by the Cell Zone Remeshing algorithm caused by having
the the prism cells in the same zone. The uniform tetrahedral mesh adapt to
shim displacement in a robust manner (no negative cells). The prism cell zone
on the seat is seen by Fluent as a stationary zone. The prism layers on the
shim is set as rigid body and assigned a dummy motion. This is a sort of
workaround in Fluent to allow the inflation layers to follow the shim boundary
and deform with it while not being part of the deforming zone. The inflation
layer motion is defined using a UDF (Appendix A) which in this case is empty.
That is no motion is assigned, yet Fluent sees the inflation layers as a moving
rigid body. This enables the feature Deform Adjacent Boundary Layer with
Zone (works only with the Spring Smoothing method).
The problem with this mesh lies in the fact that the case is started from a very
narrow gap. The inflation layers are very thin and as the gap is increased the
28 CHAPTER 3. METHOD

(a) Note the large difference in


(b) Stationary inflation layer
size between the utmost
on seat and moving inflation
inflation layer and the first
layer on shim.
tetrahedrons.

Figure 3.15: Dynamic mesh approach #4. Tet. cells and prism cells in
separate cell zones.

prism cells remain very thin compared to the rest of the mesh. The very thin
inflation layers in the gap cause a large size difference in the transition to the
tetrahedral cells which led to convergence issues. This approach could be used
with success if starting from larger gap, allowing for larger prism cells. The
ambition of this thesis is to set up an FSI case capable of simulating an entire
opening and closing cycle in the check valve, starting from a closed check valve.
Therefore this approach is rendered unsuited for the task.

Dynamic mesh approach #5 - Unstructured tetrahedral mesh of uni-


form cell size:

An unstructured tetrahedral mesh (Figure 3.16) of uniform cell size without


inflation on the boundaries was found to be the most robust alternative. It
allows the simulation to be started from an almost entirely closed gap and it
makes both local Remeshing and Cell Zone Remeshing work well.

Figure 3.16: Unstructured tetrahedral mesh without inflation layer.


Here at 0.1 mm lift.
3.3. FLUENT SETUP 29

3.3.3 Discretization

Only the swept meshes on inlet and outlet are aligned with the flow. The
unstructured tetrahedral mesh zones are not (a tetrahedral mesh is never aligned
with the flow) and in the deforming zone no boundaries have inflation layers
for good boundary layer resolution. For meshes of poor quality such as this
mesh, especially in the deforming zone, the Coupled pressure-velocity coupling
is recommended [24, p. 1464] and used for this case. Spacial discretization is
second order for all quantities to keep discretization errors to a minimum.
When remeshing and interpolating the solution onto a new mesh the temporal
discretization is first order accurate. Using only Smoothing for the dynamic
mesh would allow for a second order transient formulation. As the dynamic
mesh uses both Smoothing and Remeshing in every time step the solution will
be first order accurate in time regardless of the transient formulation setting set
in the Fluent solver input.

3.3.4 Turbulence model

The choice of turbulence model is dependent on the type of flow field to solve for
as no turbulence model is universal. In industrial CFD, two-equation turbulence
models are the most widely used [25, p. 698] and within this class of models
k  [26] and k [27] are the most common.
Both models exist in a multitude of variants. Realizable k  [28] have shown
improvements over the standard k  model where the flow exhibits strong
streamline curvature, vortices and rotation [24, p. 52] and is, in Fluent, the
recommended model from the k  family. RNG k  [29] has also shown the
same improvements over the standard model. From the k family of models
the SST k [30] is recommended in Fluent [25, p. 699].
To gauge the influence the choice of turbulence model has on the solution, a
comparative case is set up. The model is later to be compared to flow bench
measurement with flow rates up to 45 l/min. The most turbulence is expected at
the highest flow rates and this is the case used for comparison. In this operating
point the check valve will be fully open (as is seen in flow bench results presented
later). To determine whether the solution is turbulent or not levels of turbulent
kinetic energy and turbulence intensity are looked at.
The comparison is made on a mesh with inflation layers on all wall boundaries
(Figure 3.17). Although the dynamic mesh for the transient simulations will
not include inflation layers, the mesh for this comparison have been equipped
with such. The goal here is to study the difference between a turbulent vs. a
laminar solution and not the impact of grid resolution. Therefore, to make the
turbulence models perform as well as possible, the mesh is including inflation
layers. Note that with these inflation layers using the dynamic mesh approach
#4 described earlier the shim would only be able to close to a minimum gap of
about 0.4 mm. The three turbulence models are used together with Enhanced
Wall Treatment (EWT) which makes the turbulence models less sensitive to
near-wall mesh resolution. EWT uses enhanced wall functions3 in coarse near-
3 Enhanced wall functions in Fluent blends the linear wall law for the laminar viscous
30 CHAPTER 3. METHOD

wall mesh regions and in regions where the mesh is fine enough to resolve the
flow all the way down to the viscous sublayer (y + 1) a two-layer model is
used [24, p. 122].

Figure 3.17: Mesh used for turbulence model comparison.

Figure 3.18 show y + values for one of the turbulent solution. The values are in
the same range for all of the three tested turbulence models. y + is a dimension-
less wall distance defined as [31, p. 71]
u y
y+ (3.6)

where is density, y is height of first cell, is dynamic viscosity and u is
friction velocity defined as, r
w
u (3.7)

where w is wall shear stress,
 
u
w = (3.8)
y y=0

Standard log-law wall functions typically work very poorly for y + < 15. For
this mesh where y + < 3 the use of EWT is a necessity.
Figure 3.20 shows turbulent kinetic energy k plotted on a plane going through
the center of the geometry (Figure 3.19). Turbulent kinetic energy is the kinetic
energy in the modelled velocity fluctuation defined as [12, p. 563],
1 1  02 
k = u0 i u0 i = u x + u0 2y + u0 2z (3.9)
2 2
where q is the average of some quantity q. The highest levels of turbulent kinetic
energy is found in the shear layers around the jet downstream of the check valve.
Both k  models produce similar results while the k model show only half
the magnitudes of the k  models.
Velocity magnitude contours for the three turbulent solutions and the laminar
solution are shown in Figures 3.21. The flow field looks similar for all four cases
sublayer and the logarithmic wall law for the turbulent outer region using a y + dependent
blending function [24, p. 124].
3.3. FLUENT SETUP 31

Figure 3.18: Contours of y + for Realizable k  with Enhanced Wall


Treatment.

Figure 3.19: Plane for plotting solution data.

with the difference that the jet extends further downstream for the laminar
solution than for the turbulent solutions. For the three turbulent solution the
high velocity in the jet dissipates more quickly than for the laminar solution.

Reynolds number

When estimating whether a flow is laminar, turbulent or in transition, a useful


tool is the Reynolds number (Re) - a non-dimensional number relating inertial
forces to viscous forces in a fluid.
U l
Re = (3.10)

where is fluid density, is dynamic viscosity, U is characteristic velocity and
l is characteristic length [12, p. 146]. The choice of characteristic length and
velocity is dependent on the type of flow analysed.
High Re typically indicates turbulent flow while a low Re is an indication of
laminar flow. For flows inside or around more common geometries such as flow
in pipes, wide ducts or flow around airfoils, the flow can be characterized by
comparing Re to empirically found transition Re. The flow in the check valve
can not as easily be characterized. However, the flow can be split up in smaller
regions which can then be looked at individually. The flow at inlet and outlet is
32 CHAPTER 3. METHOD

(a) k - SST. (b) k  - RNG.

(c) k  - Realizable.

Figure 3.20: Turbulent kinetic energy plotted on a plane through the


domain center.

thought of as pipe flow, and characterized using pipe flow transition numbers.
For this Reynolds number study, the geometry of the dynamometer damper
is used (as shown in Figure 3.7) where the inlet and outlet of the check valve
cavity have the same cross section (14 mm in diameter). Two cases are used for
comparison - 0.1 mm lift and 1.0 mm lift with low and high flow rate.
Characteristic length for the inlet is the hydraulic diameter,

Cross-sectional area
DH 4 (3.11)
Wetted perimeter

which in the case of the inlet is,

D2 /4
DH,inlet = 4 =D (3.12)
D

where D is inlet diameter.


Characteristic velocity is the mean velocity,

Q
U= (3.13)
A
where Q is volumetric flow rate and A is cross-sectional area.
3.3. FLUENT SETUP 33

(a) k - SST. (b) k  - RNG.

(d) Without turbulence


(c) k  - Realizable.
modelling.

Figure 3.21: Contours of velocity magnitude ([m/s]) on center plane.

The flow in the gap is also characterized using the hydraulic diameter and mean
flow velocity. The hydraulic diameter of the circumferential cross section is

hDseat
DH,gap = 4 = 2h (3.14)
2Dseat

where h is lift height of the shim and Dseat is the diameter of the utmost seating
surface.
Table 3.4 lists the Reynolds number in two operating points - low lift and low
flow rate and maximum lift and high flow rate.
For flow in straight pipes, most experiments show that transition to turbulence
takes place at a Reynolds number of about Re = 3000 [12, p. 517]. If comparing
the flow at the check valve inlet and the flow in the shim-to-seat gap to straight
pipe flow, the check valve flow would appear laminar even at a high flow rate.
However, Reynolds numbers for transition to turbulent flow in straight pipes are
not directly comparable the check valve as the geometry of the check valve is
more complex. Also, most of the turbulence is likely to be produced in the shear
layers around the jet exiting the gap. Something that is not considered when
calculating the Reynolds number as done above. These difference aside, the
Reynolds numbers calculated here indicate low turbulence levels. Aside from
the turbulence produced in the jet that is.
34 CHAPTER 3. METHOD

Table 3.4: Reynolds numbers.

Operating point 1: 0.1 mm lift, p = 0.05, Q = 4.0 l/min,


Operating point 2: 1.0 mm lift, p = 0.15, Q = 41.1 l/min.

Char. length DH [m] Char. vel. U [m/s] Re


Operating point 1:
Inlet: 1.67102 0.43 104
Gap: 2.00104 10.94 108
Operating point 2:
Inlet: 1.67102 4.45 1073
Gap: 2.00103 11.24 1115

Turbulence intensity

A way to measure the level of turbulence is to relate the average fluctuating


velocity to a reference mean velocity [31, p. 64]. Turbulence intensity is defined
as
u0
T.I. (3.15)
Uref

where u0 is the root mean square of the modelled turbulent velocity fluctuations,
r   r2
0 1 02 0 2 0 2
u = ux+uy +uz = k (3.16)
3 3

Uref is chosen as the velocity at the center line of the jet (Uref = 18 m/s)
exiting the shim-seat gap as done in [32] and [33]. Figure 3.22 shows turbulence
intensity for the three turbulent solutions.
Referring to Figure 3.22, the maximum turbulence intensity is about 28% and
downstream the check valve the mean turbulence intensity is about 10%. This
indicates that the solution is in fact turbulent, especially considering the levels
are normalized with the maximum velocity. For turbulent flows like high-speed
flow inside complex geometries like for example turbines and compressor the
turbulence intensity is typically between 5% and 20% [34].
Figure 3.23 shows the volumetric flow rate across the check valve at a given
pressure difference. When comparing the flow rate between the three turbulent
solutions and the laminar solution the difference is small (about 4% between
minimum and maximum) with the laminar solution showing the lowest flow
rate for a given pressure difference.
The output parameter of interest for this study is pressure drop vs. flow rate.
Since the steady state flow rate has been shown to be rather insensitive to
whether the solution is turbulent or not, the choice is made to move forward
without a turbulence model, as the laminar solution requires less computational
resources.
3.3. FLUENT SETUP 35

(a) k - SST. (b) k  - RNG.

(c) k  - Realizable .

Figure 3.22: Contours of turbulence intensity on center plane. Refer-


ence velocity is Uref = 18 m/s measured at center of the jet exiting the
gap.

1 mm lift
50
w/o turbulence model
Exp.
Flow rate [l/min]

40 k- - SST
k- - RNG
30 k- - Realizable
Exp.
20

10

0
0.05 MPa 0.15 MPa

Figure 3.23: Volumetric flow rate at p = 0.15 MPa and 1 mm lift.

3.3.5 Mesh study

As mentioned above in Section 3.3.2 the dynamic mesh will not include inflation
layers. The forces on the shim are mostly pressure induced and any friction
induced forces are likely to be small in comparison. For this reason omitting
inflation layers, and thus sacrificing resolution in the boundary layer, might not
be too much of a simplification. It does however make the resolution in the gap
36 CHAPTER 3. METHOD

much coarser at small openings and the velocity profile in the gap might not be
accurately resolved resulting in wrong flow rates for a given pressure drop.
Below follows a brief investigation of the impact mesh resolution and the ex-
clusion of inflation layers has on the solution. Each mesh is studied in the four
operation points stated in Table 3.5. Figure 3.24 shows the meshes at 0.1 mm
lift. The meshes for the 1 mm lift case look like the one used for the turbulence
model comparison, shown in Figure 3.17.

Table 3.5: Mesh study cases.

Lift height [mm] Pressure difference [MPa]


Mesh study case 1: 0.1 0.05
Mesh study case 2: 0.1 0.10
Mesh study case 3: 1.0 0.05
Mesh study case 4: 1.0 0.15

(a) Cell size: 0.2 mm. Without (b) Cell size: 0.2 mm. With 5
inflation. 293,000 cells. inflation layers. 312,000 cells.

(c) Cell size: 0.1 mm. Without (d) Cell size: 0.1 mm. With 5
inflation. 824,000 cells. inflation layers. 861,000 cells.

Figure 3.24: Meshes used in mesh study at 0.1 mm lift height.

The resulting volumetric flow rates, shown in Figures 3.25 and 3.26 show how
at a small lift heights of the exclusion of inflation layers make a rather large
difference to the solution. At 0.1 mm lift the exclusion of inflation layers makes
for an increase in flow rate of 30-40% at the same pressure difference with these
mesh resolutions. The impact of removing the inflation layers has a greater
3.3. FLUENT SETUP 37

impact on the coarser of the two tetrahedral meshes. When comparing cell
size of the tetrahedral cells the difference in flow rate is minimal when inflation
layers are included. For the inflation-less meshes the difference in flow rate
when going from 0.1 mm to 0.2 mm cell size is 5% and 7% for 0.05 MPa and
0.10 MPa pressure difference respectively with the finer meshes as references.
It makes sense that excluding inflation layers makes a substantial difference in
solution as this small gap otherwise fits only one cell vertically with these cell
sizes.

0.1 mm lift
8
Exp. 0.2 mm tet. with infl. (312k elem.)
Flow rate [l/min]

6 0.1 mm tet. with infl. (861k elem.)


0.2 mm tet. w/o infl. (293k elem.)
Exp. 0.1 mm tet. w/o infl. (824k elem.)
4

0
0.05 MPa 0.10 MPa

Figure 3.25: Volumetric flow rate at set pressure difference and 0.1 mm
lift.

1 mm lift
50
0.2 mm tet. with infl. (312k elem.)
Exp.
Flow rate [l/min]

40 0.1 mm tet. with infl. (861k elem.)


0.2 mm tet. w/o infl. (293k elem.)
30
Exp. 0.1 mm tet. w/o infl. (824k elem.)
20

10

0
0.05 MPa 0.15 MPa

Figure 3.26: Volumetric flow rate at set pressure difference and 1 mm


lift.

The differences in solution due to mesh resolution is much smaller in the fully
open state. Here there is no trend showing. On the finer mesh, excluding the
inflation layers leads to an increase in flow rate of 2% at p = 0.05 MPa and
1% at p = 0.15 MPa. For the coarser mesh the exclusion of inflation layers
leads to a decrease in flow rate of 2% at p = 0.05 MPa. At p = 0.15 MPa
only a minimal difference in flow rate between the inflation layer equipped mesh
and the all tetrahedral mesh is observed.
Experimental data from the flow bench measurements have been added to the
bar charts for reference. These are values from the tests done on the check
valve without coil spring. It should be noted that for the CFD case, the shim
38 CHAPTER 3. METHOD

is modelled as undeformed. It is likely that the shim exhibits some bending


deformation in the experiments. At least in the 0.1 mm gap case.
When comparing the CFD results to flow bench data for the 0.1 mm lift case is
seems that excluding inflation layers brings the solution closer to the experimen-
tal data. Without inflation layers and with these cell sizes the gap is only large
enough to fit a single row of cells. Including inflation layers greatly increases
the resolution in the gap, and this should in theory make for a more accurate
solution. The fact that the inflation-less solutions correlate better to the flow
bench results should not be interpreted as excluding inflation layers produces
a more accurate solution. Since the fixed geometry CFD cases used for this
comparison does not take the bending of the shim into account, the flow rate is
under predicted and cannot be directly compared to flow bench experiments.
The above comparisons have been made on fixed geometry cases. In the coupled
case mesh resolution might have stronger influence as the mesh resolution on the
shim boundary affects the load resolution on the FSI interface. A comparison is
made on a steady state coupled case using the two tetrahedral meshes without
inflation layers. These results are presented in section 4.2.

3.3.6 Boundary conditions

For the steady state calculations the flow is driven by a pressure difference
between inlet and outlet. For model validation the same steady state pressure
drop vs. flow rate diagram produced in the flow bench is reproduced with the
FSI model. Initially the mass flow rate was set at the inlet and static pressure
set on the outlet. These boundary conditions proved too stiff. Major pressure
spikes caused large displacements of the shim leading to negative cells in the
dynamic mesh. A more robust setup is with a pressure inlet and pressure outlet.
This way the p Q diagram could be reproduced in the opposite way compared
to flow bench tests where pressure drop is measured output and flow rate is
input.
For the transient simulations where the aim is to reproduce a damper dy-
namometer test a mass flow rate inlet can successfully be used without getting
the large pressure peaks seen in the steady state case. The outlet is a pressure
outlet with static pressure set to zero as in the steady state setup. The pressure
level is that of the gas pressure, which in the dynamometer tests are 0.6 MPa,
is set under Operating Conditions in Fluent. The dynamometer test to be used
for validation is the case of a 50 Hz sinusoidal piston velocity with a maximum
amplitude of 0.25 m/s. This is represented at the inlet by a transient mass flow
rate defined in a UDF (Appendix A). Figure 3.27 shows the mass flow rate at
the inlet as a function of time. In case there is a lag in the closing phase of the
check valve the simulation time is extended a little bit beyond half the period
time to capture the entire closing phase including possible shim bounce. As the
simulation is started there is a short delay before engaging the sinusoidal BC
function. This serves as a check that the case is set up properly. If forces or
displacements show big changes in value during these initial time step there is
likely something wrong with the setup. During this initial phase the mass flow
rate is set to a small positive value.
Mass flow rate [kg/s]
3.4. STRUCTURAL SETUP
2.0 39

102

Mass flow rate [kg/s]


2.0 Sim.
end
time
0.0

0.0 2.0

0 0.5 1 1.5
Time [s] 102

Figure 3.27: Transient inlet boundary condition.

3.3.7 Fluid properties

2.0
The damper fluid is modelled in Fluent as a compressible liquid using the Tait
equation of state relating density to pressure under isothermal conditions [25,
p. 417],
 n
K0 + n(p p0 )
= (3.17)
0 K0
where is density, p is absolute pressure, K is bulk modulus and n is density
exponent. Index 0 denotes reference values at reference pressure p0 .

0
This helps reduce unphysical pressure spikes that can otherwise occur in calcu-
lations with a moving mesh. Table 3.6 presents the material data for the fluid,
and the inputs needed for the compressible liquid model in Fluent.
0.5 1
Time [
Table 3.6: Material data for the damper fluid.

Reference pressure [Pa]: 101325


Reference density [kg/m3 ]: 843
Reference bulk modulus [Pa]: 1.5109
Density exponent [-]: 1
Dynamic viscosity (constant) [kg/ms]: 0.017

3.4 Structural setup

The structural domain need only include the FSI interface and any surface that
might come in contact with it. Figure 3.28 shows the structural domain. Like
the fluid domain the structural domain is simplified by including only a 45
slice of the real geometry. The coil spring is modelled with a spring element
which is connected to the entire bottom surface of the shim as a deformable
contact, similar to a pressure load. The symmetry boundaries of the shim are
constrained in their planes by frictionless contacts.
40 CHAPTER 3. METHOD

Fixed support

Frictionless support

Fixed support

Spring element

Figure 3.28: Structural setup for the simplified case.

In the study by Leventhal [5] mentioned earlier the same spring element sim-
plification was made. He was able to successfully capture both transient effects
at check valve opening and valve bounce at check valve closure. The difference
to this case is Leventhals check valve was of ball-in-seat type while the check
valve of this study is of shim-against-seat type. The shim has some free play
radially and is thus also allowed to pitch. As a coil spring is not perfectly axial
in its reaction forces it is likely that the shim will have some radial and/or pitch
displacement. These degrees of freedom are lost when making this symmetry
simplification to the domain.
The contact between shim and seat and, in the open state, contact between
shim and spring collar are modelled with frictionless contacts with offsets to not
pinch off the fluid mesh. Details of the contact detection were set after running
standalone Structural cases with load representative of the fluid induced loads.
Values on contact detection parameters were set in a trial-and-error approach
to obtain convergence and are left to appendix (Appendix B). When running
standalone Structural cases the spring element must have a non-zero internal
damping for the solution to converge. In the coupled cases no internal damping
is used as it is assumed small in comparison to the damping of the surrounding
viscous fluid.

3.5 System Coupling setup

The procedure in System Coupling for coupling steady state cases and tran-
sients cases are similar. In the transient case multiple coupling iterations are
carried out during each coupling step (time step). In the steady state case two
approaches are possible - multiple explicit coupling steps or a single implicit cou-
pling step. The steady state result will be the same regardless of the solution
procedure as long as all quantities are converged.
3.5. SYSTEM COUPLING SETUP 41

3.5.1 Steady state

Both implicit and explicit steady solution procedures have been attempted on
this check valve case. The implicit approach with one coupling step and several
coupling iterations in conjunction with force ramping proved to give the best
control over displacement per coupling iteration. Too much displacement per
coupling iteration lead to negative cells in the dynamic fluid mesh. Using force
ramping provided some control over the displacement per time step. The fluid
forces from Fluent to Structural are then linearly ramped up over the number of
minimum coupling iterations. When reproducing the pressure vs. flow rate di-
agram from the flow bench measurements the steady state simulation is started
from a closed check valve. As the first solution is converged the boundary con-
dition is changed and the solution procedure for the second operating point is
started from previous solution. Convergence is confirmed by monitoring force
on and displacement of the shim as well as the force and displacement residuals.
When solving the steady state case in 8 operating points starting from a closed
check valve and increasing the pressure at inlet after each converged step until
the shim was fully open a minimum of 10 coupling steps was enough to make
sure the pressure was ramped up slowly enough to not cause any violent dis-
placements. The maximum number of iterations was set to 13 to let the solution
iterate a few times with full loads applied. This was enough to converge the
solution in each operating point.

3.5.2 Transient

The transient simulations replicate a damper dynamometer test with a 50 Hz


sinusoidal piston velocity. Maximum amplitude is 0.25 m/s. This experiment
show oscillations in pressure drop across the check valve with a frequency of
about 670 Hz. The ambition is to see if these pressure oscillations occur also in
the FSI solution.
Initially a time step of 1104 s is used. That is 20 time steps per the oscillation
period seen in experiments. This time step proved to long as the coupling steps
did not converge even with a as much as 10 Fluent iterations per coupling itera-
tion and 8 coupling iterations per coupling step, resulting in 80 Fluent iterations
per time step. For a transient standalone Fluent simulation the ideal number
of iterations is somewhere between 5-10 iterations per time step according to
the Fluent Manual [24, p. 1470]. If more iterations are needed to converge each
time step, the time step is too large and it is more efficient to reduce the time
step than adding to the number of iterations. For an FSI case this rule of thumb
does not directly apply as the geometry will change within the time step and
more iterations will be needed than for a standalone case but it serves at least
as a guideline indicating that the time step is to large.
With a shorter time step of 2.5105 and fewer Fluent iterations and coupling
iterations the solution converged better in the first few time step. As the shim
starts moving the solution grows unstable within the time step. These instabil-
ities are counteracted using Solution Stabilization in Fluent which slows down
the pressure response at the shim boundary.
42 CHAPTER 3. METHOD

The scale factor for the Solution Stabilization is found in a trial-and-error ap-
proach starting with no stabilization and increasing the scale factor until the
solution is stable. Note that stable in this context refers to the stability within
each time step. Figure 3.29 shows integrated pressure on the shim and dis-
placement of the shim with different scale factor values. Note that for the case
of Solution Stabilization scale factor = 0.008 the initial condition is slightly
different (lower initial mass flow rate).

2.000 103
4.128

Displacement [m]
1.500 4.126
Force [N]

1.000 4.124

0.500 4.122

0.000 4.120
0 200 400 0 200 400
Iteration Iteration
(a) Force. (b) Displacement.
104
6.000

4.000
p [Pa]

SS Scale factor = 0.000


SS Scale factor = 0.005
2.000 SS Scale factor = 0.008
SS Scale factor = 0.010
0.000
0 200 400
Iteration
(c) Pressure drop.

Figure 3.29: Tuning of Solution Stabilization scale factor. The plots


show the convergence of solution quantities within each time step during
the first time steps of the transient simulation. Iterations on the horizon-
tal axis refers to Fluent iterations. The difference in magnitude for the
case of SS Scale factor = 0.008 is due to a difference in initial conditions.

A scale factor of 0.008 is enough to stabilize the solution. Seen in Figure 3.29 is
how displacements converge nicely but the shim force does not fully converge in
these first time steps. As for the data transfer residuals, the normalized RMS
residuals for the loads sent from Fluent and displacements sent from Structural
are below 1 102 for the major part of the simulation. Figure 3.30 shows the
residuals for the load transfer from Fluent and displacement from Structural.
Another aspect to consider when deciding on time step is stability for the fluid
solution. The time stepping in Fluent is implicit and there is no stability cri-
3.6. FINAL TRANSIENT MODEL 43

Figure 3.30: Data transfer residuals.

terion to be met when deciding time step but for an efficient solution it is
recommended in Fluent that the Courant number does not exceed 20-40 in sen-
sitive transient regions [25, p. 1471]. The transient cases studied in this M.Sc.
degree project involve maximum fluid velocities of about 9 m/s s in the shim-
seat gap, which with a minimum cell size of 2104 m in the region of highest
flow velocities translates to a Courant number of about
ut 9 2.5 105
CF L = = = 2.25 (3.18)
x 2 104
which is well within the recommended limit.

3.6 Final transient model

In the final stage of the project the 45 simplification is dropped and a simu-
lation is carried out on a larger domain. As the transient simulations are to
be validated with dynamometer test the final CFD domain geometry is based
on the dynamometer damper end piece. Figure 3.7 shows the real geometry of
the damper and Figure 3.31 shows the geometry used for the fluid domain in
the final transient simulation. The hole in the center of the check valve has
been closed in the fluid domain. This is a stagnant region and the effect on the
solution is assumed negligible.
Most of the settings used in the simplified case are carried over to this larger
case with the difference that the symmetry boundary conditions are no longer
present. The mesh is larger (1.3106 cells) but the cell sizes remain the same.
The scale factor for the Solution Stabilization in Fluent is reduced to 0.002
without any instabilities showing in the solution. It might be possible to reduce
the scale factor even more but because of the long computation times for this
case these settings are not explored in-depth.
A possible explanation to the higher stability of this larger case compared to the
simplified case might be the less restrictive constraints on shim displacement.
In the simplified model the shim is constrained to the two symmetry planes,
44 CHAPTER 3. METHOD

Swept hex.
Inlet
mesh

Deforming
mesh zone

Unstructured
Outlet
tet. mesh
(a) Entire domain. (b) Mesh interior.

Figure 3.31: Fluid mesh for the final transient simulations.

while for the larger model the shim is free to move asymmetrically. As the shim
is allowed to tilt when lifting of the seat as opposed to lifting straight down in
an axisymmetric translation, the initial lift-off off the shim might be less abrupt
in the larger domain. The stability issues might also originate in the symmetry
conditions in the CFD-domain, as this is the only major difference in the fluid
model between the simplified and the final model. Aside from the size of the
domain that is.
In the structural model the spring is still modelled by a spring element, and
since the symmetry boundary condition is dropped the shim is now allowed
some radial motion and is also allowed to pitch. Radial motion is constrained
by frictionless contacts on the inner boundary with a small offset to not pinch off
the fluid mesh. Figure 3.32 shows the solid geometry and boundary conditions.

Fixed supports on
top, bottom and
innermost boundaries.

Frictionless contacts on
outer cylindrical boundary,
seat and stop.
Spring
element
Spring load applied
under the shim.

Figure 3.32: Solid domain (section view).


CHAPTER 4
Results

Below follows the results from experiment and simulation. The steady state
FSI results are compared to the quasi-steady state results from the flow bench
tests. These results are presented in the form of p Q diagram as is standard
in shock absorber analysis. The transient FSI results are compared to the
dynamometer tests. These results are presented as time series of pressure drop
and displacement, as well as solution contours on a center plane.

4.1 Flow bench measurements

The results of the flow bench measurements are presented in p Q diagrams.


Figure 4.1 shows the results from the flow bench measurements without the
spring installed for different lift heights. These results themselves are not very
interesting but they serve as a validation tool for the CFD model, since in this
test the lift height is known and can thus be replicated in a fixed geometry CFD
case. The rings mark the values used for comparison in the mesh study.
Figure 4.2 and 4.3 show the results from the measurements with the spring
installed for the two different maximum lift height. Initially the pressure drop
increases rapidly with flow rate until the pressure build-up is large enough to
overcome the spring preload, whereupon the pressure drop increases smoothly
with flow rate. In the 1 mm maximum lift case the 3 N/mm spring setup p Q
curve intersects with the no-spring setup at around 30 l/min. This is where the
preloaded shim reaches maximum lift. With the stiffer spring the shim reaches
maximum lift at a flow rate outside of the test range.
Comparing the setups with different maximum lift height (1 mm vs. 2 mm) it is
interesting to note that the pressure drop is lower for the setup with the higher
maximum lift even before the shim has reached maximum lift. From about 20
l/min and above there is a noticeable difference between the two setups with
different max. lift. Perhaps the shim opens in an asymmetrical fashion by
pitching and hitting the 1 mm stop first on one side and then continue to open
until the shim is in full contact with the stop. The flow bench is closed and it
46 CHAPTER 4. RESULTS

0.20
0.1 mm lift

Pressure drop [MPa]


0.4 mm lift
0.15 0.7 mm lift
1.0 mm lift
0.10 1.4 mm lift
2.0 mm lift

0.05

0.00
0 10 20 30 40
Flow rate [l/min]

Figure 4.1: Pressure drop at constant lift height, measured in flow


bench.

Max lift: 1 mm
0.20
w/o spring
Pressure drop [MPa]

3 N/mm, preload 3.9 N


0.15 8 N/mm, preload 10.4 N

0.10

0.05
Max. lift
0.00
0 10 20 30 40
Flow rate [l/min]

Figure 4.2: Pressure drop at 1 mm maximum lift, measured in flow


bench..

it not possible to visually monitor the shim displacement and the steady state
model is axisymmetric and thus limited to axial translation and axisymmetrical
bending. The final transient model does not have these constraints and those
results show how the pressure load is not axisymmetric and how the shim opens
by both translating and pitching. This is likely what happens in flow bench as
well, as the inlet is coming from the side like in the dynamometer damper.
4.2. STEADY STATE FSI 47

Max lift: 2 mm
0.20
w/o spring
Pressure drop [MPa]

3 N/mm, preload 3.9 N


0.15 8 N/mm, preload 10.4 N

0.10

0.05

0.00
0 10 20 30 40
Flow rate [l/min]

Figure 4.3: Pressure drop at 2 mm maximum lift, measured in flow


bench..

4.2 Steady state FSI

Figure 4.4 shows the the steady state FSI solutions for a number of operating
points, computed on the simplified domain, together with flow bench data. For
the flow bench measurements the upstream pressure is taken in the large cavity
holding the check valve fixture, where the flow is assumed stagnant. The CFD
model omits this cavity and instead the upstream pressure is taken at the inlet to
the fixture. To have comparable data the pressure drop from the FSI simulations
is the difference in total pressure at inlet and static pressure at outlet.

pf si = pin
0 p
out
(4.1)

The solution in some of the operating points have been found on two different
meshes - 0.1 mm cells and 0.2 mm cells. Both meshes are unstructured tetrahe-
dral meshes with uniform resolution and without inflation layers. The difference
is minimal.
For the case without spring where the check valve is fully open for all flow rates
the FSI model underestimates the pressure drop with 10%-20%. For the two
setups with the coil spring installed the model underestimates the pressure drop
at flow rates over 10 l/min and below that slightly overestimates the pressure
drop.
Referring to Figure 4.4, one solution point for the 8 N/mm setup stands out
(marked Discrepancy). At Q = 33 l/min the difference to the flow bench
data is 23%. At this flow rate the check valve is fully open. When solving for
the steady state solution the pressure load from the fluid field is ramped over a
number of coupling iterations. During the solve process the shim slowly moved
in small increments with each coupling iteration as the pressure was linearly
ramped up. When the shim got close to the stop the it suddenly moved in a
large increment to a fully open state. A possible explanation is that the flow in
48 CHAPTER 4. RESULTS

Flow bench, w/o spring Flow bench, 3 N/mm


Flow bench, 8 N/mm FSI, w/o spring, 0.2 mm tet.
FSI, 3 N/mm, 0.2 mm tet. FSI, 8 N/mm, 0.2 mm tet.
FSI, 3 N/mm, 0.1 mm tet. FSI, 8 N/mm, 0.1 mm tet.
0.20

0.15
Pressure drop [MPa]

0.10

Discrepancy
0.05

0.00
0 5 10 15 20 25 30 35 40 45
Flow rate [l/min]

Figure 4.4: Steady state pressure drop across the check valve vs. flow
rate. A comparison between computed results from the simplified FSI
model and flow bench measurements for three check valve setups. Max-
imum lift is 1 mm, preload for the softer 3 N/mm spring is 3.9 N and
preload of the stiffer 8 N/mm spring is 10.4 N.

the gap formed between shim and stop generates a low pressure region pulling
the shim fully open. Between the low pressure side of the shim and the shim
stop there is a small offset (20 m) to not pinch of the fluid mesh.
The steeper pressure-flow rate curve in the nose region (up to Q = 2 l/min
for the 8 N/mm setup) for the simulation compared to flow bench measurement
might be caused by the symmetry condition in the structural model. Seen in
the transient simulation is how the shim opens not by pure axial translation but
instead by pitching. The same is likely to happen in the flow bench. This might
explain why the pressure build up is not as sudden as in the FSI simulation
where the shim opens in pure axial translation.

4.3 Dynamometer tests

Figure 4.5 shows the pressure difference over the check valve measured on the
dynamometer damper, as function of time. The three tested check valve setups
show similar behaviour - an initial peak in pressure followed be oscillations at
around 670 Hz. What is interesting is that the oscillation period is the same
for the different spring rates, indicating that the oscillations are not effects
4.4. TRANSIENT FSI 49

attributed to the coil spring.

0.15 3 N/mm, preload 3.9 N


Pressure drop [MPa]

3 N/mm, preload 5.7 N


8 N/mm, preload 10.4 N
0.10

0.05
T 1.5 103 s

0.00
1 1.2 1.4 1.6 1.8 2
Time [s] 102

Figure 4.5: Dynamometer results. 50 Hz sinusoidal piston velocity with


maximum amplitude 0.25 m/s. p vs. time in rebound stroke.

4.4 Transient FSI

The same setups tested in the dynamometer are also simulated in the FSI model.
The results from the simplified model are presented in Figure 4.6. The pressure
drop is highest as the check valve first opens. The pressure drop then oscillates
before settling and then decreasing for the rest of the duration. As in the
experiments the oscillation period is the same for the softer and the stiffer
spring setup. The frequency of these oscillations are about three times as high
as those seen in the experiments.

0.10
3 N/mm, preload 3.9 N
Pressure drop [MPa]

3 N/mm, preload 5.7 N


8 N/mm, preload 10.4 N

0.05

T 0.5 103 s
0.00
0 0.2 0.4 0.6 0.8 1
Time [s] 102

Figure 4.6: FSI results from the simplified model. t = 2.5 105 s.

The larger model is tested only on one of the check valve setups - the softer
spring with the lower preload. Figure 4.7 shows pressure drop vs. time from the
simplified and the larger model as well as the experimental data. The results
50 CHAPTER 4. RESULTS

from the larger model show a higher pressure peak initially. The following
oscillations are slightly higher in frequency than those seen in the smaller model.
The biggest difference to the smaller model is the behaviour at the end of the
cycle. The results from the larger model shows a negative total pressure drop
at the end of the cycle.

Simpl. 45 domain
0.10 Full 360 domain
Pressure drop [MPa]

Dynamometer

0.05

0.00
0 0.2 0.4 0.6 0.8 1
Time [s] 102

Figure 4.7: FSI results from the simplified and the final larger model.
Spring rate: 3 N/mm, preload: 3.9 N, t = 2.5 105 s.

Figure 4.8 shows the axial displacement of the shim as function of time. Dis-
placement is measured on the nodes shown in Figure 4.9. The smaller model is
constrained to axisymmetric displacement, while in the larger model the shim is
free to move in all spatial dimensions. For the larger model the displacement is
mostly a pitch displacement. The shim opens the most on the side of the inlet.

103
0.0 Simpl. 45 domain, node 1
Axial displacement [m]

Simpl. 45 domain, node 2


Full 360 domain, node 1
Full 360 domain, node 2
Full 360 domain, node 3
0.5

1.0
0 0.2 0.4 0.6 0.8 1
Time [s] 102

Figure 4.8: FSI results from final model. Spring rate: 3 N/mm, preload:
3.9 N, t = 2.5 105 s. Shim displacement vs. time.

The largest displacement is seen at t = 0.6 102 s. Figure 4.10 shows the static
pressure on a center plane at this solution time. The inlet is coming from the
right in the figure. Figure 4.11 shows the velocity magnitude in the same plane.
Seen in both figures is how shim lifts only so slightly on the side farthest from
4.4. TRANSIENT FSI 51

(a) Simplified
domain. (b) Full domain.

Figure 4.9: Position of displacement monitors.

the inlet (to the left in the figures), and lifts almost to the stop at the inlet side.

Figure 4.10: Pressure distribution on center plane at t = 0.6 102 s


(final model).

Figure 4.11: Velocity distribution on center plane at t = 0.6 102 s


(final model).
CHAPTER 5
Conclusion and Discussion

Both steady state and transient simulations have been carried out and com-
pared to experiments. The models, both steady state and transient, have been
simplified to isolate the check valve and to make it possible to solve the cases
on a laptop computer. These simplifications include:

Reduced fluid domain.

Isothermal compressibility model.

No turbulence model.

Coil spring modelled using a spring element.

No inflation layers in the deforming mesh zone.

5.1 Conclusion

The steady state simulations show satisfactory correlation to flow bench mea-
surements at low flow rates. As flow rate increases so does the error. It should
be noted that the CFD domain does not include the entire length of piping
leading up to the downstream sensor in the flow bench, but is based on the
geometry of the check valve fixture alone. There is likely an additional pressure
drop between the outlet of the fixture and the position of the downstream pres-
sure sensor. This could explain why the FSI model predicts a lower pressure
drop than seen in the flow bench at the same flow rate.
The transient FSI model show poor correlation to dynamometer experiment.
Both the transient behaviour in the beginning of the cycle (oscillations in pres-
sure) and the amplitude of the pressure drop as the oscillations settled. This is
for both the simplified model and the final larger model. The pressure peak had
a lower amplitude in the model and the oscillations seen in the model had three
times shorter period time than those seen in the experiment. Here it is impor-
tant to note that the sampling frequency in the dynamometer experiment were
54 CHAPTER 5. CONCLUSION AND DISCUSSION

to low to resolve those frequencies. When considering the level of simplifications


made to the models, a better correlation was not expected.
It should also be noted that the dynamic effects seen in the dynamometer exper-
iments not necessarily have their origin in the check valve. In the FSI model the
check valve can easily be isolated while this is not the case in the experimental
setup. Structural compliance in the damper and friction in bushings might affect
the results. The boundary condition in CFD domain was a sinusoidal mass flow
rate to replicate the sinusoidal velocity of the dynamometer. However, look-
ing at the velocity profile of the dynamometer (Figure 3.8b) the profile is not
perfectly sinusoidal. Saying something about the accuracy of the transient FSI
model would require a larger domain including more of the damper geometry
or perhaps an experimental setup that better isolates the check valve.
The CFD mesh does not contain inflation layers on the wall boundaries in the
deforming mesh region. This to allow the shim to close to a minimal gap. It
has been shown how this simplification has an effect on the solution at small
lifts (0.1 mm), 30%-40% higher flow rate at a given pressure. For large lifts (1.0
mm) the exclusion of inflation layers had minimal effect.
For this application the choice of turbulence model has shown to be non-critical
for steady state simulations. Two of the most commonly used turbulence mod-
els have been compared to each other and to the solution without turbulence
modelling. In the flow field downstream there is a noticeable difference in how
quickly the jet dissipates. For the solution without turbulence model the jet
extends the furthest. However, this does only slightly affect the steady state
pressure difference across the check valve.
With steady state results being insensitive to turbulence modelling the choice
was made to move on to transient simulations without a turbulence model.
Although this simplification was motivated for the steady state cases the same
simplification might not be valid for transient cases, as this might lead to too
low levels of internal damping in the fluid and thus affecting dynamic behaviour
of the check valve. A turbulence model might increase the damping in the
fluid (via viscous friction losses in the modelled Reynolds stresses). For further
transient studies it is recommended to study the impact of turbulence modelling
not only on the steady state solution but the impact on transient as well.

5.2 Discussion

The goal of this M.Sc. degree project was to set up and, above all, validate an
FSI model. In doing so a number of simplifications have been made. Perhaps too
many for any model-experiment comparison to be relevant. For the steady state
validations, the CFD domain only included the fixture for mounting the check
valve in the flow bench. It would perhaps have been a better idea to include
the full length of piping in the flow bench leading up to the pressure sensors
for a better comparison. It should also be noted that the downstream pressure
sensor in the flow bench is not mounted flush with the boundary. Instead it is
connected via a junction. There might be disturbances across the junction caus-
ing significant dynamic pressures. How much this affects the pressure reading is
5.2. DISCUSSION 55

unclear. This might explain the difference in model and experiment. The flow
bench is typically used for relative measurements to compare different designs
and settings, and thus is has not been designed for absolute measurements.
Because the flow bench is run in reverse to its default setup the fixture is
mounted on top of an adapter plate that reroutes inlet and outlet piping leading
in to the flow bench main cavity. This obstruction to the flow, albeit small, be-
tween the fixture outlet and the downstream pressure sensor may give rise to a
measurable pressure drop in addition to the pressure drop over the check valve.
This might also be an explanation to the poor steady state model-experiment
correlation at high flow rates.
With steady steady state results being insensitive to turbulence modelling the
choice was made to move on to transient simulations without a turbulence
model. Although this simplification was motivated for the steady state cases
the same simplification might not be valid for transient cases, as this might lead
to too low levels of internal damping in the fluid and thus affecting dynamic
behaviour of the check valve. A turbulence model might increase the damping
in the fluid (via viscous friction losses in the modelled Reynolds stresses). For
further transient studies it is recommended to study the impact of turbulence
modelling not only on the steady state solution but the impact on transient as
well.
The reduction of the domain to a 45 slice with symmetry boundary conditions
was done to be able to solve the cases at a laptop computer. The results from
the final model, which does not have the same restrictions in degrees of free-
dom, show how the shim displacement is highly asymmetrical. The 45 slice
simplification is therefore not recommended for further studies. It did however
prove valuable to have made this simplification when exploring dynamic mesh
settings and debugging the model.
In this study the steady state FSI was a step in validating the transient model
to come. If steady state results are the goal the better practice would be to use
the dynamic mesh described in Section 3.3.2, which includes inflation layers but
prevents the shim from being able to close to a very small gap, as it would not
be necessary to start from a very small gap when just the steady state solution
is of interest.
It was assumed early on that the simulation had to be started from a very small
gap in order to capture the pressure peak at the start of the cycle. Perhaps the
transient simulation can be started from a larger gap without losing out on any
transient effects (oscillations in pressure). It would be a good idea for future
transient studies to do investigate whether starting from a larger initial gap
would affect the solution. Starting from a larger gap would allow for inflation
layers also in the deforming zone, making for better accuracy. A 20 m initial
gap was chosen as this was done by Leventhal [5] in an other check valve study.
However, the maximum lift on his check valve was smaller. In this case a larger
gap might work fine. An approach that was suggested early on was to use
a porous media to block the flow and as soon as the shim moves, remove the
porous media. This would allow for the same pressure build-up as when starting
from a very small gap.
The spring element force was applied on the shim as a distributed load on the
56 CHAPTER 5. CONCLUSION AND DISCUSSION

entire lower boundary of the shim. A more accurate representation would be


to project the upper contact surface of the coil spring onto the shim and apply
the force on that area, or better yet include the coil spring in the model as
an elastic body and drop the spring element simplification altogether. Because
of this simplified boundary condition the stresses in the shim are likely to be
far from representative of the real case, and have therefore been left out of the
report.

5.2.1 2-way FSI as a development tool

FSI can most definitely be a useful tool in shock absorber development. The
steady state model showed good correlation to experiments around opening
pressure, and can therefore be used with confidence when evaluating new design
concepts as to whether the design will work or not (open or remain closed). All
the steady state simulations in this M.Sc. degree project have been run on a
laptop computer (4 cores). This was not especially limiting as solving for one
operating point (on the smaller domain) was done in approximately 15 minutes.
Transient FSI studies require quite a lot more work compared to a steady state
case, as there is also the need to ensure convergence in every time step. The
computational requirements are typically higher for a transient simulation com-
pared to steady state. This holds true for standalone CFD cases as well as
standalone FEM cases. When coupling the two the increase in computational
requirements become even more pronounced. The transient simulations on the
smaller domain could be solved overnight ( 14 hours) on 4 cores. The final
transient simulation was solved on a 62-core cluster (Fluent using 58 cores and
structural using 4 cores) with the computation time being 4 days.
To summarize, steady state FSI is easy to set up and require small computational
resources. Transient FSI requires more work and more computational resources.

5.2.2 Suggestions for further studies

For further FSI analysis of this check valve it is recommended to use the full
geometry as computational domain. The simplified 45 domain was useful for
debugging and exploring dynamic mesh setting but the results differed from the
ones produced by the final larger model. To improve the structural model, the
coil spring could be included as an elastic body. It is however not recommended
to include the coil spring in the fluid domain. Doing so will probably make the
dynamic mesh setup troublesome. Instead a dummy body of simpler geometry
could be included in the fluid domain as a representation of the coil spring.
It is also recommended that a study on the impact of the initial gap is done be-
fore taking the transient model further. Here it was assumed that the transient
simulation must be started from a very small gap in order to get an accurate
pressure build-up at the beginning of the cycle. If this is not the case, and
the simulation can be started from a larger gap it is recommended to go with
dynamic mesh approach #4 described in Section 3.3.2, which included inflation
layers in the deforming zone.
5.2. DISCUSSION 57

For steady state simulations the simplification of not using a turbulence model to
reduce computation time has been motivated. For transient simulation however,
it is recommended to use a turbulence model because of the effects it might have
on internal damping in the fluid.
While the simplified domain was meshed with a sweep method everywhere ap-
plicable the mesh of the larger model is slightly simpler in its setup, with only
the cylindrical parts of inlet and outlet being swept. Sweepable regions can be
found also in other parts of the domain. If moving forward with this model it
could be a good idea to spend some time on this mesh to reduce the number of
elements by identifying more regions that can be swept.
For this M.Sc. degree project the pressure drop across the check valve has been
used as the solution quantity for validation. It would be a good idea to look
also at displacement. In the start of this project the ambition was to plan and
execute a dynamometer test with a transparent cavity for the check valve. Shim
displacement can then be visually monitored, and captured using a high speed
camera. This might be a suitable topic for an other degree project.
References

[1] G. E. Moore. Cramming more components onto integrated circuits. In:


Electronics 38.8 (1965), pp. 114117.
[2] N.-G. Nygren. Inside TTX - The Ohlins TTX40 Manual. 1st ed. Sweden:
Ohlins Racing AB, May 2005.
[3] Sandvik 20C for shock absorber shims. Datasheet. 2013. url: http : / /
www.smt.sandvik.com/en/materials-center/material-datasheets/
strip- steel/sandvik- 20c- for- shock- absorber- shims/ (visited on
10/13/2014).
[4] D. Borglund and D Eller. Aeroelasticity of Slender Wing Structures in
Low-speed Airflow : Lecture Notes. Stockholm, Sweden: Royal Institute of
Technology, 2013).
[5] L. Leventhal. CAE Applications in hydraulics - Experimental and analyt-
ical study of a check-valve. Tech. rep. SAE Technical Paper 2003-01-1605.
2003. doi: 10.4271/2003-01-1605.
[6] ADINA R&D, Inc. Offical webpage. url: http://www.adina.com (visited
on 10/13/2014).
[7] S. M. Price, D. R. Smith, and J. D. Tison. The effects of valve dy-
namics on reciprocating pump reliability. In: Proceedings of the Twelfth
International Pump Users Symposium. Ed. by J. C. Bailey. Texas A&M
University, Mar. 1995, pp. 221230.
[8] Exa PowerFlow. Official webpage. url: http://www.exa.com/powerflow
(visited on 11/03/2014).
[9] XFlow. Official webpage. url: http://www.xflowcfd.com (visited on
11/03/2014).
[10] C. L. Fefferman. Existence and Smoothness of the Navier-Stokes Equa-
tion. In: The Millennium Prize Problems. Ed. by J. A. Carlson, A. Jaffe,
and A. Wiles. Providence, Rhode Island: American Mathematical Society,
2006, pp. 5767. isbn: 0-8218-3679-X.
[11] Y. Du. Numerical simulation of mechanical and thermal fluid-structure
interaction in labyrinth seals. PhD thesis. TU Darmstadt, Aug. 2010.
url: http://tuprints.ulb.tu-darmstadt.de/2253/.
[12] P. K. Dowling, I. M. Kundu, and D. R. Cohen. Fluid Mechanics. Fifth
Edition. Boston: Academic Press, 2012. isbn: 978-0-12-382100-3.
60 REFERENCES

[13] COMSOL Multiphysics. Official webpage. url: http://www.comsol.se


(visited on 10/13/2014).
[14] H. Zhang and K.-J. Bathe. Direct and iterative computing of fluid flows
fully coupled with structures. In: Computational Fluid and Solid Me-
chanics. Ed. by K. J Bathe. Vol. 1. Elsevier, June 2001, pp. 14401443.
[15] S. Piperno and C. Farhat. Design of efficient partitioned procedures for
the transient solution of aeroelastic problems. In: Revue Europeenne des
Elements 9.6-7 (2000), pp. 655680. doi: 10 . 1080 / 12506559 . 2000 .
10511480.
[16] C Michler et al. The relevance of conservation for stability and accu-
racy of numerical methods for fluid-structure interaction. In: Computer
Methods in Applied Mechanics and Engineering 192.37 (2003), pp. 4195
4215.
[17] A. Dervieux et al. Total energy conservation in ALE schemes for com-
pressible flows. In: European Journal of Computational Mechanics 19.4
(2010), pp. 337363.
[18] Fluid-Structure Interaction (FSI) with ANSYS Fluent 15.0. ANSYS Cus-
tomer Portal. Training material, Lecture slides. May 2014.
[19] M. G. Doyle, S. Tavoularis, and Y. Bourgault. Application of parallel
processing to the simulation of heart mechanics. In: HPCS. Ed. by D. J.
K. Mewhort et al. Vol. 5976. Springer, 2009, pp. 3047. doi: 10.1007/978-
3-642-12659-8_3.
[20] Y. Bazilevs, K. Takizawa, and T. E. Tezduyar. Computational Fluid-
Structure Interactions: Methods and Application. Chichester, UK: Wiley,
2013. isbn: 978-0-470-97877-1.
[21] J. E. Akin, T. E. Tezduyar, and M. Ungor. Computation of flow prob-
lems with the Mixed Interface-Tracking/Interface-Capturing Technique
(MITICT). In: Computers & Fluids 36.1 (2007), pp. 211. doi: 10.1016/
j.compfluid.2005.07.008.
[22] ANSYS Fluent Dynamic Mesh (Moving Deforming Mesh) 14.5. ANSYS
Customer Portal. Training material, Lecture slides. Nov. 2013.
[23] ANSYS Meshing Users Guide. 15.0. Canonsburg, Pennsylvania: ANSYS,
Inc., Nov. 2013.
[24] ANSYS Fluent Theory Guide. 15.0. Canonsburg, Pennsylvania: ANSYS,
Inc., Nov. 2013.
[25] ANSYS Fluent Users Guide. 15.0. Canonsburg, Pennsylvania: ANSYS,
Inc., Nov. 2013.
[26] B. Launder and D. Spalding. The numerical computation of turbulent
flows. In: Computer Methods in Applied Mechanics and Engineering 3.2
(1974), pp. 269 289. doi: 10.1016/0045-7825(74)90029-2.
[27] D. C. Wilcox. Formulation of the k- turbulence model revisited. In:
AIAA Journal 46.11 (2008), pp. 28232838.
[28] T.-H. Shih et al. A new k- eddy viscosity model for high Reynolds num-
ber turbulent flows - Model development and validation. In: Computers
& Fluids 34.3 (1995), pp. 227238. doi: 0.1016/0045-7930(94)00032-T.
REFERENCES 61

[29] S. A. Orszag et al. Renormalization Group Modeling and Turbulence


Simulations. In: Near-wall turbuent flows. Ed. by R. M. C. So, C. G.
Speziale, and B. E. Launder. Amsterdam, Netherlands: Elsevier, 1993,
pp. 10311046.
[30] F. R. Menter. Two-equation eddy-viscosity turbulence models for engi-
neering applications. In: AIAA Journal 32.8 (1994), pp. 15981605. doi:
10.2514/3.12149.
[31] H. K. Versteeg and W. Malalasekera. An Introduction to Computational
Fluid Dynamics: The Finite Volume Method. Second Edition. Pearson
Publications, 2007. isbn: 978-81-317-2048-6.
[32] A. Hashiehbaf and G. P. Romano. Experimental investigation on circu-
lar and non-circular synthetic jets issuing from sharp edge orifices. In:
Proceedings of 17th International Symposium on Applications of Laser
Techniques to Fluid Mechanics. Lisbon, Portugal, July 2014.
[33] C. Bogey and C. Bailly. A study of the influence of the Reynolds number
on jet self-similarity using large-eddy simulation. In: Direct and Large-
Eddy Simulation VII. Ed. by V. Armenio, B. Geurts, and J. Frohlich.
Vol. 13. Springer Netherlands, 2010, pp. 1116. isbn: 978-90-481-3651-3.
doi: 10.1007/978-90-481-3652-0_2.
[34] CFD Online - Turbulence intensity. Webpage. 2012. url: http://www.
cfd-online.com/Wiki/Turbulence_intensity (visited on 10/13/2014).
APPENDIX A
UDFs and Profiles

Dummy UDF for inflation layers in separate


cell zone
#include "udf.h"

DEFINE_GRID_MOTION(dummy_motion, domain, dt, time, dtime)


{
return;
}

Profile for previewing dynamic mesh motion

(
(open_n_close_valve point 4 1)
(time 0.0000 0.0095 0.0096 1.0000)
(v_y -0.1 -0.1 0.1 0.1)
)
64 APPENDIX A. UDFS AND PROFILES

UDF for transient inlet BC (360 case)


#include "udf.h"
#define PI 3.14159265359

static real time_prev = 0.0;

static real mfr_in_0 = 0.0001; /* Initial flow rate [kg/s] */


static real mfr_in_max = 0.21452; /* Maximum flow rate [kg/s] */
static real freq = 50; /* Frequency [Hz] */
static real time_init = 0.0003; /* Delay to let flow field initialize [s] */

DEFINE_PROFILE(mfr_in_t,thread,pos)
{
#if !RP_HOST

face_t f;
real time, dtime, mfr_in_ampl, mfr_in;

time = CURRENT_TIME;
dtime = CURRENT_TIMESTEP;
mfr_in_ampl = mfr_in_max - mfr_in_0;

/* Update mass flow inlet at each time step */


if(fabs(time_prev-time) > 0.2*dtime)
{
/* Current mass flow inlet */
if(time <= time_init)
{
mfr_in = mfr_in_0;
}
else
{
mfr_in = mfr_in_ampl * sin(2*PI*freq * (time-time_init)) + mfr_in_0;
}

/* Loop over the inlet faces */


begin_f_loop(f,thread)
{
F_PROFILE(f,thread,pos) = mfr_in;
}
end_f_loop(f,thread)

/* print new mass flow rate to TUI*/


Message0("\n\n ************* MASS FLOW INLET INFO *************\n");
Message0("\n Mass flow rate : %5.3e kg/s", mfr_in);
Message0("\n Flow time : %5.3e s", time);
Message0("\n\n ************************************************\n\n");

/* assign current time */


time_prev = time;
}
#endif
}
APPENDIX B
Model settings (360 case)

Fluent setup

Floating point format . . . . . Double precision

Solution Setup

Solver
Type . . . . . . . . . . . . . . . . . . . . . . . . Pressure-Based
Time . . . . . . . . . . . . . . . . . . . . . . . . Transient
Models
Energy. . . . . . . . . . . . . . . . . . . . . . . Off
Viscous Model . . . . . . . . . . . . . . Laminar
Materials
Fluid . . . . . . . . . . . . . . . . . . . . . . . . Damper fluid (refer to Table 3.6)
Boundary conditions
Inlet . . . . . . . . . . . . . . . . . . . . . . . . mass-flow-inlet, (refer to Figure 3.27)
Outlet . . . . . . . . . . . . . . . . . . . . . . pressure-outlet, 0 Pa
Operating Pressure . . . . . . . . . . 1106 Pa
Dynamic Mesh
Mesh Methods . . . . . . . . . . . . . . Smoothing, Remeshing
Smoothing
Method Diffusion
Diffusion Function Boundary Distance
Diffusion Parameter 0
66 APPENDIX B. MODEL SETTINGS (360 CASE)

Remeshing
Remeshing Methods Local Cell, Local Face
Min. Length Scale 6105 m
Max. Length Scale 2.5104 m
Max Cell Skewness 0.95
Size Remeshing Interval 1
Solution Stabilization
Method . . . . . . . . . . . . . . . . . . . . . . . coefficient-based
Scale Factor . . . . . . . . . . . . . . . . . . . 0.002

Solution

Pressure-Velocity Coupling
Scheme . . . . . . . . . . . . . . . . . . . . . Coupled
Spatial Discretization
Pressure . . . . . . . . . . . . . . . . . . . . . Second Order
Density . . . . . . . . . . . . . . . . . . . . . Second Order Upwind
Momentum . . . . . . . . . . . . . . . . . Second Order Upwind
Transient Formulation
Scheme . . . . . . . . . . . . . . . . . . . . . First Order Implicit
Solution Controls
Flow Courant Number . . . . . . 20
Explicit Relaxation Factors . . Momentum: 0.75
Pressure: 0.75
Time Step
Max Iterations/Time Step 4

Structural setup

Connections

Contacts:

Definition
Type . . . . . . . . . . . . . . . . . . . . . . . . Frictionless
Advanced
Formulation . . . . . . . . . . . . . . . . . Augmented Lagrange
Normal Stiffness . . . . . . . . . . . . Manual
Normal Stiffness Factor . . . . . 0.5
Pinball Region . . . . . . . . . . . . . . Radius
Pinball Radius . . . . . . . . . . . . . . 2.5103 m
Time Step Controls . . . . . . . . . None
APPENDIX B. MODEL SETTINGS (360 CASE) 67

Geometric Modification
Interface Treatment . . . . . . . . . Add Offset, No Ramping
Offset . . . . . . . . . . . . . . . . . . . . . . . 2.0105 m (seat and stop),
1.0104 m (inner boundary)

Spring element:

Definition
Type . . . . . . . . . . . . . . . . . . . . . . . . Longitudinal
Longitudinal Stiffness . . . . . . . 3000 N/m
Preload . . . . . . . . . . . . . . . . . . . . . Free Length
Free Length . . . . . . . . . . . . . . . . . 6.5103 m
Scope
Scope . . . . . . . . . . . . . . . . . . . . . . . Body-Ground
Mobile
Behavior . . . . . . . . . . . . . . . . . . . . Deformable

Mesh

Body Sizing - Shim:

Definition
Type . . . . . . . . . . . . . . . . . . . . . . . . Element Size
Element Size . . . . . . . . . . . . . . . . 3.0104 m
Behavior . . . . . . . . . . . . . . . . . . . . Soft

Analysis Settings

Step Controls
Auto Time Stepping . . . . . . . . On
Define By . . . . . . . . . . . . . . . . . . . Substeps
Initial Substeps . . . . . . . . . . . . . 1
Minimum Substeps . . . . . . . . . . 1
Maximum Substeps . . . . . . . . . 5
Time Integration . . . . . . . . . . . . On
Solver Controls
Solver Type . . . . . . . . . . . . . . . . . Iterative
Large Deflections . . . . . . . . . . . . On
68 APPENDIX B. MODEL SETTINGS (360 CASE)

System Coupling Setup

Analysis Settings

Initialization Controls
Coupling Initialization . . . . . . Program Controlled
Duration Controls
Duration Defined By . . . . . . . . End Time
End Time . . . . . . . . . . . . . . . . . . . 0.011 s
Step Controls
Step Size . . . . . . . . . . . . . . . . . . . . 2.5105 s
Minimum Iterations . . . . . . . . . 2
Maximum Iterations . . . . . . . . 5

Execution Control

Co-Sim. Sequence
Transient Structural . . . . . . . . . 1
Fluent . . . . . . . . . . . . . . . . . . . . . . 2

Das könnte Ihnen auch gefallen