• 沒有找到結果。

B Collapseofaliquidcolumn:numericalsimulationandexperimentalvalidation

N/A
N/A
Protected

Academic year: 2022

Share "B Collapseofaliquidcolumn:numericalsimulationandexperimentalvalidation"

Copied!
25
0
0

加載中.... (立即查看全文)

全文

(1)

Comput. Mech. (2007) 39: 453–476 DOI 10.1007/s00466-006-0043-z

O R I G I NA L PA P E R

Marcela A. Cruchaga · Diego J. Celentano Tayfun E. Tezduyar

Collapse of a liquid column: numerical simulation and experimental validation

Received: 19 December 2005 / Accepted: 11 January 2006 / Published online: 11 February 2006

© Springer-Verlag 2006

Abstract This paper is focused on the numerical and exper- imental analyses of the collapse of a liquid column. The measurements of the interface position in a set of experi- ments carried out with shampoo and water for two different initial column aspect ratios are presented together with the corresponding numerical predictions. The experimental pro- cedure was found to provide acceptable recurrence in the observation of the interface evolution. Basic models describ- ing some of the relevant physical aspects, e.g. wall friction and turbulence, are included in the simulations. Numerical experiments are conducted to evaluate the influence of the parameters involved in the modeling by comparing the re- sults with the data from the measurements. The numerical predictions reasonably describe the physical trends.

Keywords Moving interfaces· Two-fluid flows ·

Computational fluid mechanics· Experimental validation 1 Introduction

Modeling two-liquid interfaces or free-surface flows is an active research field due to the wide range of their engineer- ing applications. This motivates the development of formula- tions capable of accurately representing this kind of transient problems. Various numerical techniques were proposed, with moving or fixed domain discretizations, to overcome the chal- lenges related to updating the interface intrinsically coupled with the fluid dynamics equations (e.g. see [1–34], and refer- ences therein). Work was carried out also in the development of physical models to represent real problems at laboratory scales and, in particular, to assess the numerical performance M.A. Cruchaga (

B

)· D.J. Celentano

Departamento de Ingeniería Mecánica, Universidad de Santiago de Chile, Av. Bdo. O’Higgins 3363, Santiago, Chile E-mail: mcruchag@lauca.usach.cl T.E. Tezduyar

Team for Advanced Flow Simulation and Modeling (T*AFSM), Mechanical Engineering, Rice University - MS 321,

Houston, TX 77005, USA

of the proposed numerical techniques by comparing their pre- dictions with the measurements obtained from experiments.

Focusing on the description of the collapse of a water col- umn, experiments were presented in [11, 35]. This problem was adopted by several researchers as a benchmark test to validate the numerical performance of the proposed formu- lations in solving two-liquid interfaces or free-surface flows [11, 12, 15, 21, 23, 24, 26, 28, 29, 32].

In this paper, in addition to numerical results, we report results from experiments for the collapse of liquid columns.

The experiments are carried out using two different fluids:

shampoo and water, and two different column aspect ratios (height to width ratio of the initial liquid column). The pur- pose of these experiments is to provide data for the interface evolution with very different fluid properties. The experi- mental data is used for evaluating the performance of the numerical formulations we propose for solving this class of problems.

We use finite element techniques with fixed meshes. The interface is “captured” by solving an advection equation gov- erning the time-evolution of an interface function that marks the interface location [16–33]. A version of this approach presented in [32] can capture the interface while maintain- ing the global mass conservation and accurately representing the sudden changes in the materials properties. That formu- lation is called the edge-tracked interface locator technique (ETILT), which was introduced in [19]. More recent versions of the ETILT were introduced in [30, 31, 33]. Although some new techniques related to the ETILT are presented in this work, the simulations are mainly devoted to evaluation of the numerical performance of our core techniques in describing some of the physical aspects such as wall friction and turbu- lence. The influence of the parameters involved in modeling is also analyzed.

The governing equations are presented in Sect. 2. The ETILT and the new aspects included in its current version are summarized in Sect. 3. The details of the experimental procedure are presented in Sect. 4. Experimental and numeri- cal results are presented and discussed in Sect. 5. Concluding remarks are presented in Sect. 6.

(2)

454 M.A. Cruchaga et al.

Fig. 1 Collapse of a liquid column: experimental setup and illustrative picture

2 Governing equations

The Navier–Stokes equations of unsteady incompressible flows are written as follow:

ρ∂u

∂t + ρu∇ · u + ∇ p − ∇ · (2με) = ρf in  × Y, (1)

∇ · u = 0 in  × Y, (2)

whereρ, u, p, μ, ε and f are the density, velocity, pressure, dynamic viscosity, strain-rate tensor, and the specific body force. In these equations, denotes an open-bounded domain with a smooth boundary, and Y is the time interval of inter- est. This system of equations is completed with a set of initial and boundary conditions:

u= u0 in (3)

u= g in g× Y (4)

σ · n = h in h× Y (5)

where u0 is the initial value of the velocity field, g repre- sents the velocity imposed on the part of the boundaryg, and h is the traction vector imposed overh (g∪ h =  andg∩ h = ∅), typically taken as traction-free condition:

h= 0.

In the present simulations, the condition at a friction wall is given by the following simple model [21]:

h= −ρu |u|

C2f in(h)w× Y, (6)

where Cf is a friction parameter and(h)w is a subset of

h, representing the walls where the friction conditions are applicable.

A simple model to compute the energy dissipated by the turbulent effects could be obtained by modifyingμ as [32]:

μ = min(μ + lmix2 ρ

2ε : ε; μmax), (7)

where lmix is a characteristic mixing length (in the present work, lmix = CthUGNwith Ct being a modeling parameter and hUGNa characteristic element length [20]) andμmaxis a cut-off value.

(3)

Collapse of a liquid column 455

Fig. 2 Collapse of a liquid column for aspect ratio Ar= 1. Experimental results for the time evolution of the interface position: a dimensionless horizontal position (x/L) at y = 0.0 m (bottom wall) and b dimensionless vertical position (y/L) at x = 0.0 m (left wall). Comparison with the results obtained by Martin and Moyce (M&M) [35]

The interface between the two fluids (Fluids 1 and 2) rep- resents a strong discontinuity in the fluid properties and the gradients of the velocity and pressure. Nevertheless, these variables are interpolated as continuous functions across the interface. Other types of discontinuities at the interface, e.g.

surface tension, are not included in the present model. The interface motion is governed by an advection equation:

∂ϕ

∂t + u · ∇ϕ = 0 in  × Y, (8)

whereϕ is a function marking the location of the interface.

The weak form of the Navier–Stokes equations is obtained in the context of a finite element formulation using a Gen- eralized Streamline Operator technique [34]. The stabilized nature of the technique allows the use of equal-order inter- polation functions for velocity and pressure. In the context of two-fluid flow analysis, all the matrices and vectors de- rived from the discrete form are computed including the dis- continuity in fluid properties [32]. Therefore, the algorithm used for updating the interface is coupled with the finite element solution of the Navier–Stokes equations. The time integration is performed using a standard backward Euler scheme.

(4)

456 M.A. Cruchaga et al.

Fig. 3 Collapse of a liquid column for aspect ratio Ar= 2. Experimental results for the time evolution of the interface position: a dimensionless horizontal position (x/L) at y = 0.0 m (bottom wall) and b dimensionless vertical position (y/L) at x = 0.0 m (left wall). Comparison with the results obtained by Martin and Moyce (M&M) [35] and Koshizuka and Oka (K&O) [11]

3 Interface update

In the present work, the version of the ETILT presented in [32] is used for updating the interface. At each time step, Eqs. (1) and (2) are computed in the entire domain using the density and viscosity distributions expressed as:

ρh = ϕheρ1+ (1 − ϕhe2, (9.1)

μh= ϕheμ1+ (1 − ϕhe2, (9.2) whereϕheis the edge-based representation ofϕ.

To determine uhn+1, phn+1andϕnhe+1at a time level n+ 1 from unh andϕnhe at time level n, the velocity and pressure are computed from Eqs. (1) and (2). To computeϕnhe+1, given ϕhen , first a nodal representationϕhis computed. This is done by using a constrained least-squares projection as given in [30, 33]:

(5)

Collapse of a liquid column 457

Fig. 4 Collapse of a shampoo column with aspect ratio Ar = 1. Interface positions at different instants: experimental (left) and numerical results computed without (Simulation 1, middle) and with (Simulation 2, right) wall friction effects





ψh

ϕnh− ϕnhe

 d

+

nie



k=1

ψh(xkPEN

ϕnh(xk) − 0.5

= 0. (10)

Hereψh is the test function, nieis the number of the inter- face edges (i.e., the edges crossed by the interface), xkis the

coordinate of the interface location along the kth interface edge andλPEN is a penalty parameter. This penalty param- eter ensures that the projection topologically preserves the interface position along edges. In four-noded quadrilateral elements, this condition is activated also along the element diagonals crossed by the interface. After this projection, we update the interface by using a discretization based on the streamline upwind/Petrov–Galerkin (SUPG) formulation [1]

of the advection equation (8) governingϕ:

(6)

458 M.A. Cruchaga et al.

Fig. 5 Collapse of a shampoo column with aspect ratio Ar = 1: a dimensionless horizontal position (x/L) at y = 0.0 m (bottom wall), b dimensionless vertical position (y/L) at x = 0.0 m (left wall), c dimensionless vertical position (y/L) at x = 0.27 m (middle of the box) and d dimensionless vertical position (y/L) x = 0.42 m (right wall). Numerical results computed without (Simulation 1) and with (Simulation 2) wall friction effects





ψh

∂ϕh

∂t + uh.∇ϕh

 d

+

nel



e=1



e

τSUPGuh.∇ψh ∂ϕh

∂t + uh.∇ϕh

 d

+

nel



e=1



e

∇ψhνDCID∇ϕhd = 0 (11)

Here nelis the number of elements,τSUPGis the SUPG sta- bilization parameter [20], andνDCIDis the discontinuity-cap- turing interface dissipation (DCID) proposed as:

νDCID = Cs

2 h2UGN

2ε : ε∇ϕhhUGN

ϕref , (12)

where Csis a discontinuity-capturing parameter andϕrefis a reference value set to 1. We note that the justification behind the expression given by Eq. (12) is combining some of the features we see in Eq. (7) and the discontinuity-capturing

(7)

Collapse of a liquid column 459

Fig. 5 (Contd.)

directional dissipation (DCDD) given in [20].ϕnh+1is com- puted from Eq. (11) by using a Crank–Nicholson time inte- gration scheme.

Fromϕnh+1we obtainϕnhe+1by a combination of a least- squares projection:





ψnhe+1



P

nhe+1)P− ϕnh+1

d = 0 (13)

and corrections to enforce volume conservation for chunks of Fluids 1 and 2 [32]. The subscript P in Eq. (13) is used for representing the intermediate values following the projec- tion, but prior to the corrections for the volume conservation.

In this technique,nhe+1)P = H

ϕnh+1− 0.5

, where H is the Heaviside function. The volume conservation condition is given by the following equation:





ϕnhe+1− ϕnhe



d = Q, (14)

where Q is the mass inflow/outflow in the time interval [n, n + 1]. An iterative procedure is employed to satisfy Eq.

(14). The mass balance ratio is defined as Rm =

⎣



nhe+1− ϕnhe)d

⎦

Q for Q= 0, (15)

(8)

460 M.A. Cruchaga et al.

Fig. 6 Collapse of a shampoo column with aspect ratio Ar= 2. Interface positions at different instants: experimental (left) and numerical results computed without (Simulation 1, middle) and with (Simulation 2, right) wall friction effects

Rm =

⎝



ϕnhe+1d

⎠ 

ϕnhed



for Q= 0. (16)

The residual Rm needs to be equal to 1.0 for volume con- servation. To achieve this, ϕnh+1 is corrected iteratively as follows:

ϕhn+1,i+1= ϕnh+

ϕnh+1,i− ϕnh

/ (Rm)k, (17)

where k= sign

ϕnh+1,i− ϕnh



and i is the iteration counter.

The iterations continue until the volume conservation condi-

tion is reached, i.e.|Rm− 1.0| < εR, whereεRis an accept- able tolerance. When the convergence is reached,ϕhn+1gen- erates a volume-conserving value forϕnhe+1. This value serves as the starting point for the computations in marching from n+ 1 to n + 2.

4 Experimental procedure

The collapse of a water column was extensively studied in open channels [35] and in closed containers [11]. In the

(9)

Collapse of a liquid column 461

Fig. 6 (Contd.)

present work, a simple experimental setup was built to com- pare the experimental data with the numerical results to be evaluated in Sect. 5. The apparatus used is shown in Fig. 1 and consists of a glass box with two parts delimited by a gate with a mechanical release system. The left part of the box is initially filled with a liquid: shampoo or coloured water.

In both cases two different liquid aspects ratios were tested:

Ar = 1 and Ar = 2 (Ar = H/L, with H and L being the initial height and width of the column respectively, with L = 0.114 m).

Different experiments were carried out and recorded to a film. To measure the interface position, a ruler with tics 0.02 m apart is drawn surrounding the box and over the frame of the gate. A red marker that jointly moves with the gate helps to observe its position during the gate opening. The measurements reported in the present work are average val- ues from the measurements registered for a minimum of five experiments. The maximum experimental errors in space and time are estimated as ±0.005 m and ±0.025 s, respectively.

(10)

462 M.A. Cruchaga et al.

Fig. 7 Collapse of a shampoo column with aspect ratio Ar = 2: a dimensionless horizontal position (x/L) at y = 0.0 m (bottom wall), b dimensionless vertical position (y/L) at x = 0.0 m (left wall), c dimensionless vertical position (y/L) at x = 0.27 m (middle of the box) and d dimensionless vertical position (y/L) x = 0.42 m (right wall). Numerical results computed without (Simulation 1) and with (Simulation 2) wall friction effects. Simulation 3 was carried out with the wall friction effects and a finite gate opening speed

To evaluate the performance of the experimental proce- dure, the observed evolution of the interface at the bottom and left walls of the box are compared in Figs. 2 and 3 with those reported in [11] and [35], up to the instant that the liquid impinges on the right wall. The dimensionless hori- zontal position (x/L) at the bottom wall versus dimension- less time (t[Arg/L]1/2, with t being the real time and g the gravity modulus) is plotted in Figs. 2a and 3a for the two

aspect ratios and both liquids. Figures 2b and 3b show for the same cases the evolution of the dimensionless vertical position (y/L) at the left wall (time scaled as in Figs. 2a and 3a). Figures 2 and 3 show that the measurements are in rea- sonably good agreement with those reported in the literature (in particular for water with Ar = 2). Note that the inter- face positions are delayed in time for the shampoo due to its higher viscosity.

(11)

Collapse of a liquid column 463

Fig. 7 (Contd.)

5 Comparison of numerical and experimental results In [32] we discussed some performance aspects of the ETILT:

sensitivity to mesh and time-step sizes, performance for struc- tured and unstructured meshes, quadrilateral and triangu- lar finite element discretizations, different sets of properties for the two liquids, and laminar and turbulent flow simu- lations. Also, in that a work the collapse of a water col- umn was analyzed with a low value of the initial column width (L = 0.05715 m). The computed results, up to the instant of the water impingement on the right wall, were

compared with those obtained using a Lagrangian technique and other numerical techniques and measurements reported in the literature.

In the present paper, the numerical simulations are focused on studying the long-term transient behavior for the collapse of a liquid column described in Sect. 4. In particular, the ef- fects of the parameters used in modeling some of the physical aspects (wall friction and turbulence) as proposed in Sect. 3 are evaluated by comparing them with the experimental re- sults. The influence of the DCID is also analyzed. In the sim- ulations, the liquid column is initially at rest and confined

(12)

464 M.A. Cruchaga et al.

Fig. 8 Collapse of a shampoo column with aspect ratio Ar= 2. Interface positions at different instants: effect of the gate opening

Fig. 9 Collapse of a shampoo column with aspect ratio Ar= 2. Interface positions at different instants: 3D analysis

between the left wall and the gate. The pressure is set to zero at the top of the rectangular computational domain.

5.1 Collapse of a shampoo column

The fluid properties are:ρ1=1, 042 kg/m3andμ1=8 kg/m/s for the shampoo (measured at a laboratory using a standard procedure), andρ2= 1 kg/m3andμ2= 0.001 kg/m/s for the air (lower values for the air viscosity do not play a significant role in the numerical predictions, as observed in [32]).

As the shampoo-glass contact effects are observed in experiments, simulations with slip (Simulation 1) and fric-

tion (Simulation 2) conditions at the walls were carried out to assess the influence of this effect on the interface response.

In Simulation 2, we set Cf = 190 [see Eq. (6)] based on the numerical tests described below. A mesh composed of 60× 45 four-noded isoparametric elements is used with a time-step size of 0.001 s.

Column of aspect ratio Ar = 1 Figure 4 shows the exper- imental interface positions at various instants together with the corresponding predictions computed via Simulations 1 and 2. A detailed description of the interface evolution is shown in Fig. 5.

(13)

Collapse of a liquid column 465

Fig. 10 Collapse of a water column with aspect ratio Ar= 1. Interface positions at different instants: experimental (left) and numerical results computed without (Simulation 1, middle) and with (Simulation 2, right) discontinuity-capturing dissipation

The interface evolution at the bottom of the box dur- ing the spread of the liquid under the gravity is plotted in Fig. 5a up to the instant when the shampoo interface reaches the right wall (t ≈ 0.4 s). The motion of the horizontal position of the interface is delayed due to the wall friction effects.

The time histories for the vertical position at the left wall, middle of the box and right wall are shown in Figs. 5b, c and d. From these figures, it is apparent that, in the experiment, the

shampoo motion practically stops once the interface impinges on the right wall. Although similar predictions of the inter- face vertical position at the left wall are obtained for Simula- tions 1 and 2, some differences with the experimental results persist. As can be seen, the maximum interface vertical po- sition at the middle of the box is captured reasonably well in the simulations, but the results exhibit a time delay. With the friction wall model, the maximum value of the interface vertical position at the right wall is reduced. Nevertheless,

(14)

466 M.A. Cruchaga et al.

Fig. 10 (Contd.)

Fig. 5 shows that the overall trends for the interface behavior are captured reasonably well.

Column of aspect ratio Ar = 2 Figure 6 shows the experi- mental interface positions at different instants and the numer- ical results obtained with Simulations 1 and 2.

To evaluate the effect of the gate opening, an additional computation is carried out with the gate not opening instan- taneously but with a finite speed (Simulation 3). The gate opening speed was extracted from the experiments. In this simulation, an average gate opening speed of 0.86 m/s is used,

while the friction conditions are used on both the walls and the gate.

Evolutions of the horizontal position of the interface at the bottom of the box and the vertical position at the left wall, middle of the box and right wall are shown in Fig. 7 together with the experimental results.

The simulations show that the interface, as it was also observed in the experiments, still moves after the first impingement on the right wall (t ≈ 0.3 s). The reflected wave realistically represents the real liquid motion. The results computed with Simulation 2 satisfactorily capture the

(15)

Collapse of a liquid column 467

Fig. 10 (Contd.)

maximum and minimum interface vertical positions and the time of the first impingement on the right wall.

The gate opening effect included in Simulation 3 slightly changes the results obtained in Simulation 2. With the gate opening effect, more advanced interface positions at early instants of the analysis are obtained at the bottom of the box.

This is due to the “orifice effect” induced at the beginning of the gate opening, which increases the velocity at the bot- tom of the box. This situation quickly changes and the liquid

attaches to the right wall later than it does in Simulation 2. The interface positions at different instants obtained with Simulation 3 are shown in Fig. 8.

In the present study, numerical analyses with different values of the friction coefficient Cf were carried out for the shampoo columns with aspect ratios 1 and 2 (results not shown). The value Cf = 190 was selected on the basis that it leads to minimum differences between the numerical predictions of the interface position and the corresponding

(16)

468 M.A. Cruchaga et al.

Fig. 11 Collapse of a water column with aspect ratio Ar = 1: a dimensionless horizontal position (x/L) at y = 0.0 m (bottom wall), b dimen- sionless vertical position (y/L) at x = 0.0 m (left wall), c dimensionless vertical position (y/L) at x = 0.27 m (middle of the box) and d dimensionless vertical position (y/L) at x = 0.42 m (right wall). Numerical results computed without (Simulation 1) and with (Simulation 2) discontinuity-capturing dissipation

measurements. Such differences can not be fully eliminated with just a friction model, because other aspects, such as sur- face tension (not included in the model), should also play a role.

In the computations reported above, the shampoo was considered a Newtonian fluid. A more realistic representa- tion of the material behavior for the shampoo can be based on choosing a viscosity that decreases with the increasing shear rate. To evaluate this effect, an additional computation was carried out using a simple constitutive

model based on the power law of the shear rate. The pre- liminary results (not shown) did not show any quantitative changes.

We also carried out a 3D computation based on the con- ditions of Simulation 1. These results are included to briefly assess the performance of the proposed technique in a 3D case. The interface position is shown in Fig. 9, and we see that the interface evolution is close to what we obtained in the 2D analysis. This trend agrees with the fact that 3D effects were not observed in the experiments.

(17)

Collapse of a liquid column 469

Fig. 11 (Contd.)

5.2 Collapse of a water column

In this case the gate is assumed to be suddenly removed at time t = 0 s. Slip conditions are assumed at the solid sur- faces. The fluid properties are:ρ1 = 1, 000 kg/m3and μ1

= 0.001 kg/m/s for the water, and ρ2 = 1 kg/m3 and μ2

= 0.001 kg/m/s for the air.

Two simulations are carried out: Simulations 1 and 2.

Both simulations include a simple turbulence model (defined by Eq. (7) and with Ct = 3.57). Simulation 2 also has the DCID (defined by Eq. (12) and with Cs = 10). The value

of the coefficient Cs was selected from numerical tests by trying to minimize the differences between the numerical pre- dictions and the experimental measurements for both aspect ratios.

The mesh is composed of 100× 75 four-noded isopara- metric elements. The time-step size is 0.001 s.

Column of aspect ratio Ar= 1 The experimental and numer- ical interface positions at various instants are shown in Fig. 10.

The cut-off value of the viscosity in Eq. (7) is set as μmax = 1.5 kg/m/s. In the experiments, it is observed that

(18)

470 M.A. Cruchaga et al.

Fig. 12 Collapse of a water column with aspect ratio Ar = 2. Interface positions at different instants: experimental (left) and numerical results computed without (Simulation 1, middle) and with (Simulation 2, right) discontinuity-capturing dissipation

separated flows are developed close to the interface at specific instants (e.g., during the liquid impingement on the walls).

The separated fluid quickly merges with the bulk liquid and no bubbles are seen. However, bubbles are obtained in Sim- ulation 1 for t ≥ 0.6 s. Although they are caused due to the folding of the interface, they have no physical meaning be- yond that. Note that the formulation does not include a bubble model. When the DCID is activated (Simulation 2), the for- mation of bubbles in the bulk liquid are inhibited and the interface is represented as a continuous front.

Figure 11 shows, for the experiments and Simulations 1 and 2, the evolutions of the interface horizontal position at the bottom of the box and the interface vertical position at the left wall, middle of the box and right wall. The results are in reasonably good agreement with the experimental data.

The computed vertical position of the interface at the left wall practically coincides with the experimental results up to t = 0.9 s. Additionally, the magnitude of its second maxi- mum is well described. The numerical behavior of the inter- face vertical position at the middle of the box also shows good

(19)

Collapse of a liquid column 471

Fig. 12 (Contd.)

agreement between the numerical and experimental results up to t = 0.6 s. After that, we see some discrepancies in the maximum values and a delay in the occurrence of the sec- ond maximum. The computed vertical interface position at the right wall captures reasonably well the instant of the first impingement (t = 0.25 s) and first maximum (t = 0.45 s).

The maximum value, however, is lower than that observed experimentally. The second maximum is described reason- ably well in magnitude but has a delayed time.

Column of aspect ratio Ar=2 In this case μmax=3.0 kg/m/s.

Figure 12 shows the interface positions at various instants.

The time history of the interface horizontal position along the bottom of the box and the interface vertical position at selected locations are shown in Fig. 13.

The numerical results obtained in this case show trends that are similar to those observed in the previous analy- sis. In particular, the numerical evolutions of the interface positions at the bottom of the box and along the left wall

(20)

472 M.A. Cruchaga et al.

Fig. 12 (Contd.)

satisfactorily match the corresponding measurements up to t = 1.5 s. Moreover, the time of the first impingement on the right wall is captured well. The amplitude of the reflected wave is predicted reasonably well but has a delayed time.

6 Conclusions

The performance of the ETILT was assessed in simulation of the collapse of liquid columns. A set of experiments have

been carried out to obtain the evolution of the interface posi- tion. Two different fluids, shampoo and water, and two differ- ent column aspect ratios, 1 and 2, were studied.

Simple concepts were included in the model to describe the wall friction and turbulence effects in the analyses of the shampoo and water, respectively. Additionally, the influ- ence of some proposed stabilization parameters was assessed.

The wall friction effect is particularly apparent in the collapse of the shampoo column due to its high viscosity.

(21)

Collapse of a liquid column 473

Fig. 12 (Contd.)

No separated flow was observed in the experiments. This aspect was also seen in the simulations where neither tur- bulence model nor discontinuity-capturing dissipation was used. The effect of the gate opening was found not to change the overall behavior of the interface. In addition, the 3D anal- ysis provides results similar to those computed under the 2D assumption.

In the case of the collapse of the water column, since it is practically an inviscid fluid, the friction wall effect does

not play much role in the analyses. No special model was in- cluded to represent the bubbles. Hence, once the bubbles are generated, the interface becomes unrealistic. The disconti- nuity-capturing dissipation included in the formulation helps to model the overall interface behavior reasonably well and inhibits the bubble formation in the bulk liquid. The turbu- lence model controls the height of the liquid columns on the walls and the wave evolution. A high mesh refinement is required to better represent the turbulence effects, and

(22)

474 M.A. Cruchaga et al.

Fig. 13 Collapse of a water column with aspect ratio Ar = 2: a dimensionless horizontal position (x/L) at y = 0.0 m (bottom wall), b dimen- sionless vertical position (y/L) at x = 0.0 m (left wall), c dimensionless vertical position (y/L) at x = 0.27 m (middle of the box) and d dimensionless vertical position (y/L) at x = 0.42 m (right wall). Numerical results computed without (Simulation 1) and with (Simulation 2) discontinuity-capturing dissipation

reasonable results are obtained with the coefficients used in the turbulence model.

Overall, the numerical results compare satisfactorily with the measurements.

Acknowledgements The support provided by the Chilean Council of Research and Technology CONICYT (FONDECYT Projects No 1020029 and 7020029) and the Department of Technological and Scien- tific Research at the University of Santiago de Chile (DICYT-USACH) is gratefully acknowledged.

(23)

Collapse of a liquid column 475

Fig. 13 (Contd.)

References

1. Brooks AN, Hughes TJR (1982) Streamline upwind/Petrov–Galer- kin formulations for convection dominated flows with particular emphasis on the incompressible Navier–Stokes equations. Com- put Methods Appl Mech Eng 32:199–259

2. Hughes TJR, Liu WK, Zimmermann TK (1981) Lagrangian-Eule- rian finite element formulation for incompressible viscous flows.

Comput Methods Appl Mech Eng 29:239–349

3. Liu WK (1981) Finite element procedures for fluid-structure inter- actions and application to liquid storage tanks. Nucl Eng Des 65:221–238

4. Huerta A, Liu W (1988) Viscous flow with large free surface motion. Comput Methods Appl Mech Eng 69:277–324

5. Tezduyar TE (1992) Stabilized finite element formulations for incompressible flow computations. Adv Appl Mech 28:1–44 6. Tezduyar TE, Behr M, Liu J (1992) A new strategy for finite ele-

ment computations involving moving boundaries and interfaces – the deforming-spatial-domain/space-time procedure: I. The con- cept and the preliminary numerical tests. Comput Methods Appl Mech Eng 94:339–351

7. Tezduyar TE, Behr M, Mittal S, Liu J (1992) A new strategy for finite element computations involving moving boundaries and interfaces – the deforming-spatial-domain/space-time procedure:

II. Computation of free-surfaces flows, two-liquid flows, and flows

(24)

476 M.A. Cruchaga et al.

with drifting cylinders. Comput Methods Appl Mech Eng 94:

353–371

8. Braess H, Wriggers P (2000) Arbitrary Lagrangian–Eulerian finite element analysis of free surface flow. Comput Methods Appl Mech Eng 190:95–109

9. Feng YT, Peri´c D (2003) A spatially adaptive linear space-time finite element solution procedure for incompressible flows with moving domains. Int J Numer Methods Fluids 43:1099–1106 10. Rabier S, Medale M (2003) Computation of free surface flows with

a projection FEM in a moving mesh framework. Comput Methods Appl Mech Eng 192:4703–4721

11. Koshizuka S, Oka Y (1996) Moving-particle semi-implicit method for fragmentation of incompressible fluid. Nucl Sci Eng 123:421–

434

12. Bonet J, Kulasegaram S, Rodriguez-Paz MX, Profit M (2004) Var- iational formulation for the smooth particle hydrodynamics (SPH) simulation of fluid and solid problems. Comput Methods Appl Mech Eng 193:928–948

13. Kulasegaram S, Bonet J, Lewis RW, Profit M (2004) A variational formulation based contact algorithm for rigid boundaries in two- dimensional SPH applications. Comput Mech 33:316–325 14. Xie H, Koshizuka S, Oka Y (2004) Modelling of a single drop

impact onto liquid film using particle method. Int J Numer Meth- ods Fluids 45:1009–1023

15. Idelsohn S, Storti M, Oñate E (2003) A Lagrangian meshless finite element method applied to fluid-structure interaction problems.

Comput Struct 81:655–671

16. Tezduyar T, Aliabadi S, Behr M (1998) Enhanced-discretiza- tion interface-capturing technique (EDICT) for computation of unsteady flows with interfaces. Comput Methods Appl Mech Eng 155:235–248

17. Osher S, Fedkiw P (2001) Level set methods: and overview and some recent results. J Comput Phys 169:463–502

18. Sethian JA (2001) Evolution, implementation, and application of level set and fast marching methods for advancing fronts. J Comput Phys 169:503–555

19. Tezduyar TE (2001) Finite element methods for flow problems with moving boundaries and interfaces. Arch Comput Methods Eng 8:83–130

20. Tezduyar TE (2003) Computation of moving boundaries and inter- faces and stabilization parameters. Int J Numer Methods Fluids 43:555–575

21. Kim MS, Lee WI (2003) A new VOF-based numerical scheme for the simulation of fluid flow with free surface. Part I: New free sur- face-tracking algorithm and its verification. Int J Numer Methods Fluids 42:765–790

22. Minev P, Chen T, Nandakumar K (2003) A finite element technique for multifluid incompressible flow using Eulerian grids. J Comput Phys 187:255–273

23. Sochnikov V, Efrima S (2003) Level set calculations of the evo- lution of boundaries on a dynamically adaptive grid. Int J Numer Methods Eng 56:1913–1929

24. Yue W, Lin CL, Patel VC (2003) Numerical simulation of unsteady multidimensional free surface motons by level set method. Int J Numer Methods Fluids 42:853–884

25. Tezduyar TE, Sathe S (2004) Enhanced-discretization space-time technique (EDSTT). Comput Methods Appl Mech Eng 193:1385–

1401

26. Kohno H, Tanahashi T (2004) Numerical analysis of moving inter- faces using a level set method coupled with adaptive mesh refine- ment. Int J Numer Methods Fluids 45:921–944

27. Wang JP, Borthwick AGL, Taylor RE (2004) Finite-volume-type VOF method on dynamically adaptive quadtree grids. Int J Numer Methods Fluids 45:485–508

28. Greaves D (2004) Simulation of interface and free surface flows in a viscous fluid using adapting quadtree grids. Int J Numer Methods Fluids 44:1093–1117

29. Cruchaga MA, Celentano DJ, Tezduyar TE (2004) Modeling of moving interface problems with the ETILT. In: Computational mechanics proceedings of the WCCM VI in conjunction with AP- COM’04. Tsinghua University Press and Spring-Verlag, Beijing, China

30. Tezduyar TE (2004) Finite elements methods for fluid dynamics with moving boundaries and interfaces. In: Stein E, De Borts R, Hughes TJR (eds) Encyclopedia of computational mechanics, Flu- ids, vol 3, chapt 17. Wiley, New York

31. Tezduyar TE (2004) Moving boundaries and interfaces. In: Franca LP, Tezduyar TE, Masud A (eds) Finite element methods: 1970’s and beyond. CIMNE, Barcelona, pp 205–220

32. Cruchaga MA, Celentano DJ, Tezduyar TE (2005) Moving-inter- face computations with the edge-tracked interface locator tech- nique (ETILT). Int J Numer Methods Fluids 47:451–469 33. Tezduyar TE (2006) Interface-tracking and interface-capturing

techniques for finite element computation of moving boundaries and interfaces. Comput Methods Appl Mech Eng (published online)

34. Cruchaga MA, Oñate E (1999) A generalized streamline finite element approach for the analysis of incompressible flow prob- lems including moving surfaces. Comput Methods Appl Mech Eng 173:241–255

35. Martin J, Moyce W (1952) An experimental study of the collapse of liquid columns on a rigid horizontal plane. Philos Trans R Soc Lond 244:312–324

(25)

參考文獻

相關文件

Nonreciprocal Phenomena in Chiral Materials - Left and Right in Quantum Dynamics –..

• LQCD calculation of the neutron EDM for 2+1 flavors ,→ simulation at various pion masses &amp; lattice volumes. ,→ working with an imaginary θ [th’y assumed to be analytic at θ

1: Invariant mass distribution and the leading lepton transverse momentum in data and simulation..

Examples of thermal image (left) and processed binary images (middle and right) of

These images are the results of relighting the synthesized target object under Lambertian model (left column) and Phong model (right column) with different light directions ....

• measuring 1st qubit gives 2 possible results. – 0 with the probability 1/2, and the post-measurement

Robert G.Sargent ,「 Validation and Verification of Simulation Models」 ,Proceeding of the 1992 Winter Simulation

This study, analysis of numerical simulation software Flovent within five transient flow field,explore the different design of large high-temperature thermostat room