1. Background and Objectives
Modern finite element methods (FEM) and computational fluid dynamics (CFD) programs have enabled engineers to extensively use simulation tools in the design and manufacturing stages [
1,
2]. As a result, more complex fluid and solid models are directly utilized in the design and analysis phase for a better understanding of various complex engineering systems. For the steady delivery of internal fluids, major and minor friction losses provide design engineers with rough estimates of required pressure drops for a range of volume flow rates; moreover, specific in-depth studies are still needed for specific components of the delivery systems, for instance, the needle valve and the hoses with thermal protective layers [
3,
4,
5]. Moreover, the study of the initial transients with respect to different rheological properties and fluid–structure interaction effects is very much needed [
6,
7]. For this type of transient viscous fluid with extremely high dynamic viscosities, the complex rheological properties play more significant roles than convection, along with turbulent effects [
8,
9]. In this paper, through further studies of the complex non-Newtonian polymers utilized in EV manufacturing plants and their respective delivery systems, the importance of effective hierarchical computational models is demonstrated in the comparison of analytical, experimental, and computational models [
10,
11].
Earlier studies of the controlled dispensing of non-Newtonian glues utilized in EV manufacturing plants are presented in Refs. [
12,
13]. As shown in
Figure 1, a bucket of the non-Newtonian glue within the dispense system is hydraulically actuated. Because of the extreme dynamic viscosity of the non-Newtonian fluid, very high pressure within the range of a few hundred times that of the atmospheric pressure must be introduced for intermittent dispensing. Moreover, due to the glue’s temperature dependence regarding rheological properties, the T-shape dispensing gun connects the reservoir through hoses with extensive thermal protective layers. Moreover, the needle valve within the gun is actuated by a hydraulic piston [
5]. In this study, we focus on the interaction between the non-Newtonian, nearly compressible fluid with flexible tubes and hierarchical modeling strategies similar to that of complex structures for more physical insights about complex fluid dispensing systems [
11]. Similar to the study on water hammering, we demonstrate that a weak shock, the so-called pressure wave or acoustic signals, can also be carried through the fluid, which is assumed to be an almost compressible or nearly incompressible medium [
5,
14], yet the added mass, added stiffness, and the viscosity, in this case, provide a fluid delivery system with excessive overall damping. As reported in Refs. [
12,
13,
15], numerous tests have confirmed that given a level of inlet and outlet pressure drop, the flow rate or the average axial velocity tends to approach the steady solution in a way similar to that of a resistor and inductor/capacitor (RL/RC) circuits or Kelvin/Maxwell viscoelastic behaviors. Moreover, because of the extreme viscosity, the glue dispensing through the T-shape gun, the servo-actuated needle valve, and flexible tubes exhibits similar initial transient behaviors. Hence, the precision control of the dispensing system for such non-Newtonian, nearly incompressible fluids demands a better understanding of the developing fluid characteristics. Finally, inverse optimization methods were implemented for the accurate predictions of one, two, or three system parameters.
2. Complex Rheological Properties
In EV manufacturing plants, various glues are utilized for sealing and other applications. The complex non-Newtonian rheological properties have presented tremendous challenges in the precision control of dispensing systems, in particular, with the considerations of variable speeds, durations, and flow rates [
16,
17]. In our earlier studies of such non-Newtonian fluid models, a Bingham non-Newtonian model with a peak dynamic viscosity of
a minimum dynamic viscosity of
and a critical shear rate of
has been employed. The rheology of the polymer material in question is very much shear- and strain- rate-dependent [
18,
19]. In the computational studies presented in this paper, we employ both the power law and thixotropic non-Newtonian, almost compressible, or nearly incompressible, rheological models as depicted in
Figure 2.
In engineering design, it is common to relate the shear stress
with the shear rate
. There are, in general, two types of non-Newtonian fluids, namely, shear thinning and shear thickening, based on the tabulated relationship between the dynamic viscosity
and the shear rate
, utilizing the definition of the dynamic viscosity
for fluid. The stress tensor
can be expressed as
, where the kinematic viscosity is defined as
with the density
, and the shear strain rate tensor
can be denoted as
. For the shear-thinning non-Newtonian fluids of interest to us, we employ a power law model,
, in which the equivalent dynamic viscosity
can be expressed as
where
A and
a, which are also expressed as
in some of the literature, are constants, and
is the effective deformation rate or shear rate defined as
Based on the experimental data presented in Ref. [
13], we have identified the constant
and the power
. In this study, we also adopt a pragmatic approach with a simple thixotropic, almost compressible non-Newtonian fluid model, as presented in Equation (
2), which exhibits an initial dynamic viscosity
with an initial relaxation time
and a final dynamic viscosity
with a final relaxation time
Note that in Equation (
2), key parameters can be fine tuned or optimized with the inverse optimization procedure in Refs. [
12,
20]. This important data analysis approach can be very effective in linking computational, analytical, and experimental evidence directly with needed recommendations for design and manufacturing. Furthermore, based on experimental evidence of black glue, we employ
. The final dynamic viscosity level
is adjusted. For the pipe diameter
for the volume flow rate
, the average axial velocity
is around
, the equivalent shear rate is
, and the corresponding initial dynamic viscosity
is about
, whereas the final dynamic viscosity
is about
As confirmed in
Figure 2, within the shear rate according to the operation conditions, the thixotropic material does exhibit the so-called shear thinning effects for high shear rates, as well as the so-called shear thickening effects towards extremely high shear rates. In this paper, in addition to the thixotropic non-Newtonian properties illustrated in Equation (
2), we employ the density
and the compressibility
Notice that the compressibility
, which is also the reciprocal of the bulk modulus
, is defined as
or
, with the volume
V, the specific volume
v, or the reciprocal of the density
, and the pressure is
p. Assume that the initial pressure and density for this newly proposed thixotropic non-Newtonian, almost compressible model are
and
, such that we have
where the wave speed
c is assumed to be constant, defined as
3. Dry Structure Vibrations
Due to the excessively high pressure utilized in the dispensing of the thixotropic non-Newtonian, almost compressible fluid in EV manufacturing plants, which, in some cases, can exceed
or
the thickness-to-diameter ratio of the tube is more than
. For instance, in the example elaborated in this study, the wall thickness is
, and the tube diameter is merely
. Therefore, it is necessary to introduce the so-called thick-wall cylindrical pressure vessel model. Consider a differential element, as shown in
Figure 3; we have the force equilibrium in the tangential or
direction,
along with the force equilibrium in the radial or
r direction,
where we have
and
Moreover, we can simplify the governing equation in both tangential and radial directions as
where
and
stand for accelerations in the radial and tangential directions.
As a special case, within an axial directional cross-section, for the static or quasi-static force equilibrium with
and
, we have
In fact, Equation (
7) implies an axisymmetric nature, namely, no dependence on the polar angle
. Hence, we have the definition of the so-called hoop strain
, a normal strain in the tangential or
direction, defined as
where
u stands for the radial direction expansion displacement, which is only dependent on the radial position marked by the radius
r. Likewise, we can define the radial normal strain
as
.
For linear elastic materials, Hook’s law yields
where
E and
stand for Young’s modulus and Poisson’s ratio.
The linear system equations in Equation (
8) yield
Thus, in combining Equations (
7) and (
9), we derive the following Cauchy–Euler equation with respect to the radial displacement
,
It is very important to observe the consistency of the units in Equation (
10), which implies the characteristic solutions in the form of an algebraic or power function
. If we insert the trial solution
, we have the following characteristic equation:
Hence, we have the two characteristic solutions to Equation (
10), namely,
and
, and as a consequence,
with
and
.
Finally, the normal stresses
and
within the
z-direction cross-section can be simply expressed as
With the internal and external pressure boundary conditions, we have
at
and
at
; we derive
Using Equation (
8), we have the following radial direction displacement distribution:
where the two constants
and
are expressed as
In the physical models presented in this paper, the pipe diameter
D is
, the pipe length
L is
, and the pipe thickness
d is
. The length-to-diameter ratio
is 20; thus, it is different from the immediate neighborhood around both end supports. The radial displacements of these dry structures should resemble those illustrated in
Figure 4 and
Figure 5.
Figure 4 depicts a comparison between a finite element solution of a long structure and an infinitely long, thick-wall cylindrical pressure vessel theoretical solution. The results other than the immediate neighborhood of the fixed ends compare well within 4 to 9 percentage points. Furthermore, as shown in
Figure 5, the dry structure response, in which a steady inlet-to-outlet pressure distribution is applied to the interior of the tube, matches well with the quasi-static expansion of the flexible tube. In this paper, we assume the same Poisson’s ratio of
for both steel and aluminum, whereas the Young’s modulus for the aluminum is set at
and
for steel. Moreover, in the example chosen for this study, the length of the tube is
, which is 20 times the diameter. Thus, other than the inlet and outlet structural supports with
, the radial expansion characterized by the displacement
depends on the axial position via the linear pressure distribution from the inlet to the outlet. As shown in
Figure 6, the radial direction displacement for the aluminum tube is nearly three times that of the steel tube, just as their respective Young’s modulus suggests. Moreover, the dynamic response of the flexible tube, as depicted in
Figure 6, indicates that the response frequency
f is around
, namely, the natural frequency
is around
which will be compared with the estimate of the acoustic frequencies.
4. Acoustic Ranges
In this section, we focus on the respective acoustic ranges of the thick-wall tube. Unlike electromagnetic waves, which can propagate through a vacuum, sound, as a form of mechanical vibration, must have a compressible continuum for its waves to propagate, with its energy carried by partly kinematic and partly potential energies [
21]. The key assumption of the transmission of sound in air or any other continuum is to ignore the shear effects [
22,
23]. Consequently, the mass conservation is represented by the relationship between the pressure
p and the volumetric strain
, where
represents the displacement vector of the medium. Thus, for any acoustic medium, we have
where
stands for the body force, and the stress tensor
can be simply expressed as
.
In practice, for solids, the bulk modules
can be expressed as
, a function of Young’s modulus
E and Poisson’s ratio
. It is important to recognize that acoustic waves do not trigger large motions or displacements. Hence, derived from Newton’s Second Law of Motion, we have
Furthermore, we also have the pressure and the volumetric strain relationship
Employing Equations (
15) and (
16), we derive the following key equations:
where the wave speed
c is expressed as
with the bulk modulus
and the density
.
For the cylindrical pipe, as illustrated in
Figure 7, it is convenient to express Equation (
17) in the cylindrical coordinate system. Using the separation of variables method, we have
By substituting the pressure expression into the governing equation in Equation (
17), we can derive
For the
z-direction or the axial direction with the two fixed surfaces at
and
, the special solution can be expressed as
With respect to the
r-direction or the radial direction, we have a constant pressure or, rather, zero gauge pressure at
and
. Notice that the steady solution with two different pressures on the internal and external cylindrical surface, namely,
and
, will be combined with the transient solutions, for which the gauge pressures at
and
are zero, as depicted in
Figure 7. Consequently, the normal gradient of the displacement, velocity, and acceleration is zero based on Equation (
16). For the polar angle
, a periodic function with a period of
is assigned; therefore, we have
where
and
are constants dependent on boundary conditions.
Therefore, when combining the special functions in the axial or
z and the polar or
directions, we have
with the modes
m and
n in
and
z directions, respectively.
Finally, the solution in the radial or
r direction is governed by the following Bessel function
with
Note that the solution to Equation (
19) is a combination of a typical Bessel function of the first kind of order
m, namely
, and a typical Bessel function of the second kind of order
m, namely,
. In this dispensing system, because of the extremely high pressure introduced to overcome the excessive viscosity, the tube or pipe resembles a typical thick wall cylindrical pressure vessel with
Thus, the special solution can be expressed as a combination of a Bessel function of the first kind
and a Bessel function of the second kind
. From Equation (
18), we have
, where
and
are constants that are dependent on the initial and boundary conditions, and the natural frequency
will be determined by the boundary conditions of
, namely
. Notice that nontrivial solutions for
and
must exist for
We must have the determinant of the coefficient matrix equal to zero, namely
As depicted in
Figure 8, with respect to the mode number
, we have the first group of characteristic solutions
as
,
,
and
, respectively, and the second group of characteristic solutions
as
,
,
, and
. Notice that for the geometries selected in this study, the separations of these acoustic frequency ranges are not significant. Furthermore, using Equation (
20), we have
where the wave number is defined as
with the wavelength
Thus, we have the first four lowest natural frequencies
and
listed as follows:
Finally, we have the dominant wave length listed as
For the aluminum pipe, the bulk modulus
is roughly
, whereas for the steel pipe, the bulk modulus
is nearly
In this work, we also adopt a density for aluminum at
, whereas the density for steel is
Hence, the wave speeds
c for the aluminum and steel tubes are
and
respectively, which confirms the simulation results depicted in
Figure 6, namely, the dry structure wave length for the aluminum tube is nearly the same as that of the steel tube. Finally, by employing Equation (
24), we can estimate the acoustic range natural frequency
to be around
or the frequency
f to be around
for the aluminum tube, and we can estimate
or the frequency
f to be around
for the steel tube. Based on the transient responses of both dry structures, as depicted in
Figure 6, the structural frequency
f is around
It is clear that, in comparison, the acoustic frequencies are much higher than the structural frequencies; we focused our attention on structural vibrations for the full-fledged fluid–structure interaction (FSI) analyses.
5. FSI Effects and Developing Flow Transients
In the study of the dispensing of thixotropic non-Newtonian, almost compressible fluids, we have established that a significant inlet and outlet pressure difference is needed to squeeze out the super viscous polymers [
13]. With the consideration of initial transients, namely, the inlet flow rate is elevated to the steady level within a short period, in this paper, we adopt a thixotropic model, as depicted in
Figure 2. In the series of FSI simulations using Bentley Systems ADINA FSI software, we employed a time step size of
, along with a ramp size of
As demonstrated in
Figure 9,
Figure 10,
Figure 11,
Figure 12 and
Figure 13, the dry structure exhibits much less damping in comparison with FSI systems with internal viscous fluids. Moreover, in so-called dry structural models, the same pressure distribution from the inlet to the outlet established in the steady solution tends to introduce a linear structural dynamic response that satisfies the structural boundary conditions, which are fixed at both ends. Because the high aspect ratio is
, for the bulk of the tube, the radial direction displacements still resemble those from the thick-wall cylindrical pressure vessel, as shown in
Figure 4,
Figure 5 and
Figure 6.
It is clear that in our Bentley Systems ADINA FSI models, which are closer to the physical realities, the structural deformation will follow a dynamic response from the inlet to the outlet, yet with significantly more damping in comparison with their respective dry structures, namely, with the damping effects of the very viscous thixotropic non-Newtonian, almost compressible internal fluid. As discussed in the previous section, since the aluminum density-to-Young’s modulus ratio is nearly the same as the steel density-to-Young’s modulus, even though the aluminum Young’s modulus is about one third of that of steel, the wave number
k seems to be the same for both aluminum and steel pipes, as confirmed in
Figure 11 and
Figure 13. Since the wave propagation speeds for both aluminum and steel pipes are similar, the natural frequencies of the dry structures are also the same, as confirmed in
Figure 6. Clearly, due to the damping of the viscous fluid, the transient structural solutions will eventually approach those of dry structures, as shown in
Figure 6,
Figure 11 and
Figure 13.
For the precise delivery of these very complicated thixotropic non-Newtonian, almost compressible fluids, the wave speed, which is dependent on the bulk modulus or the reciprocal of the compressibility, as well as the fluid–structure interaction (FSI) effects, can play an important role. Moreover, as shown in
Figure 14 and
Figure 15, although both compressible fluid models and full-fledged FSI models yield oscillatory pressure differences and average axial velocity during the developing flow phase, the frequencies have significantly increased, whereas the damping ratios stay roughly the same. It is not difficult to conclude that the added stiffness part of the FSI effects seems to be more prominent, which matches with our physical intuitions, namely, both aluminum and steel tubes are very strong, and the rigid tube assumption is reasonable. Moreover, as shown in
Figure 14 and
Figure 15, both FSI and CFD analyses produce accurate solutions oscillating around the steady theoretical solutions, as elaborated in our earlier reports and Refs. [
12,
13].
Furthermore, as demonstrated in
Figure 14 and
Figure 15, comparing the pressure and the average velocity for the CFD case using rigid tubes with those of deformable tubes, it is clear that the slight deformation of the pipe is actually very conducive to the axial motion of the internal fluid. In fact, for biological systems such as the cardiovascular and digestive systems, the motion of the tube wall plays a significant role in conveying fluids and solids. This is an interesting point that is worth noting from a design and manufacturing perspective. In addition to the variabilities of dynamic viscosity, it is also important to estimate the effects of the pipe length for these non-Newtonian fluids, a process similar to the establishment of major friction loss. For the cases presented in
Figure 14 and
Figure 15, the flow rate is
or
which has an average axial velocity of
for the pipe with a diameter of
Thus, it takes about
for the Lagrangian fluid particle to move from the inlet to the outlet. Overall, the full-fledged FSI models are depicted in
Figure 9 and
Figure 10. To further digest these results, the dry structure dynamic results are also compared with the FSI results in
Figure 11 and
Figure 13. It is clear that the internal fluid contributed significantly to the damping effects.
As shown in
Figure 14, the oscillation of both the average axial velocity and the pressure difference between the inlet and the outlet around their respective steady state values tends to dissipate and diminish within the first few milliseconds of the initial transients. This time scale provides important guidelines for EV manufacturing dispensing systems with respect to how fast the activation and closing of the servo motor and the needle valve are. To further the study on the compressibility effect within dispensing systems, we can also investigate the pressure wave and the velocity propagation within the rigid tube. Notice that the Dirichlet pressure boundary at the outlet corresponds to the Neumann velocity boundary with a zero-velocity axial derivative, whereas the Dirichlet velocity boundary at the inlet corresponds to the Neuman pressure boundary with a zero-pressure axial derivative, as elaborated in Refs. [
12,
13].
Finally, as suggested in
Figure 15, for a given flow rate, the inlet and outlet pressure drop for such non-Newtonian internal fluids over time behaves very much like a mechanical vibration system with a critical damping ratio. To characterize the system parameters, such as relaxation time, based on the experimental or computational results, in this study, we employed the so-called inverse optimization method with Newton–Raphson iterative procedures. A good validation of the inverse optimization method is the viscoelastic dynamic model of a downhole sucker rod and its system parameters, which yield almost identical results in comparison with the respective experimental measures as reported in Ref. [
12]. The key viscosity effects are damping for both the inlet and outlet pressure drop and the axial velocity transients, as illustrated in
Figure 11 and
Figure 13. Again, as the dynamic viscosity reaches a critical level, both the inlet and outlet pressure difference and the axial velocity will approach steady solutions, following a similar exponential growth pattern as reported in Ref. [
13]. This information is crucial for the design of control algorithms for the dispensing system of many different types of complex fluids.