16.11.2014 Views

Mixing Process in a Water Storage Tank by CFD Simulations

Mixing Process in a Water Storage Tank by CFD Simulations

Mixing Process in a Water Storage Tank by CFD Simulations

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Mix<strong>in</strong>g</strong> <strong>Process</strong> <strong>in</strong> a <strong>Water</strong> <strong>Storage</strong> <strong>Tank</strong> <strong>by</strong> <strong>CFD</strong> <strong>Simulations</strong><br />

I.J. Moncho-Esteve and G. Palau-Salvador P.A. López-Jiménez<br />

Department of Rural Eng<strong>in</strong>eer<strong>in</strong>g Hydraulic and Environmental<br />

Eng<strong>in</strong>eer<strong>in</strong>g Department<br />

Polytechnic University of Valencia<br />

Polytechnic University of Valencia<br />

Valencia<br />

Valencia<br />

Spa<strong>in</strong><br />

Spa<strong>in</strong><br />

W. Brevis M.O. Vaas<br />

Department of Civil and Structural Eng<strong>in</strong>eer<strong>in</strong>g Institute for Hydromechanics<br />

The University of Sheffield<br />

Karlsruhe Institute of Technology<br />

Sheffield<br />

Karlsruhe<br />

UK<br />

Germany<br />

Keywords: <strong>CFD</strong>, Turbulent <strong>Mix<strong>in</strong>g</strong>, <strong>Water</strong> <strong>Storage</strong> <strong>Tank</strong>, <strong>Mix<strong>in</strong>g</strong> Parameters.<br />

Abstract<br />

<strong>Mix<strong>in</strong>g</strong> processes and their related residence times <strong>in</strong> water storage tanks are<br />

an important factor for ma<strong>in</strong>ta<strong>in</strong><strong>in</strong>g an adequate quality of water. These processes<br />

are normally <strong>in</strong>vestigated with costly experimental studies. Furthermore, much has<br />

been achieved <strong>in</strong> numerical simulation techniques <strong>in</strong> recent years. The ma<strong>in</strong><br />

objective of this project is to develop a <strong>CFD</strong> (Computational Fluid Dynamics) model<br />

that could be used as a flexible modell<strong>in</strong>g tool, <strong>in</strong> addition with experimental<br />

procedures, for the study, design and improvement of these tanks. <strong>Mix<strong>in</strong>g</strong> and<br />

residence times <strong>in</strong> a water storage tank are <strong>in</strong>vestigated us<strong>in</strong>g Unsteady Reynolds<br />

Averaged Navier-Stokes simulations (URANS). A good agreement between the<br />

experimental dilution curve <strong>in</strong> the outlet section tank and the <strong>CFD</strong> results was<br />

achieved. The flow structure and its relation to the different phases of the dilution<br />

curve have been studied, prov<strong>in</strong>g a predom<strong>in</strong>antly advective transport.<br />

INTRODUCTION<br />

<strong>Water</strong> storage tanks are important devices <strong>in</strong> hydraulic networks. They are<br />

designed for several objectives such as keep<strong>in</strong>g water supply constant when water<br />

demand achieves flow peaks. Although the study and modell<strong>in</strong>g of these tanks are often<br />

simplified, the impact of the tanks for ma<strong>in</strong>ta<strong>in</strong><strong>in</strong>g an adequate quality of water <strong>in</strong> the<br />

network is of paramount importance. Some examples of fatal episodes due to <strong>in</strong>adequate<br />

management are documented <strong>by</strong> Bolous et al. (1994) and Clark et al. (1993).<br />

<strong>CFD</strong> methods can be seen as the best analysis and predictive tool <strong>in</strong> storage tanks<br />

for water quality studies (Mart<strong>in</strong>ez-Solano et al., 2010), s<strong>in</strong>ce these models can obta<strong>in</strong><br />

sufficient spatial and temporary detail of the studied doma<strong>in</strong>s. However, storage tanks,<br />

have been normally studied with <strong>in</strong>tensive experimental procedures (Patwardhan, 2002),<br />

<strong>in</strong> which it is assumed an empirical or statistical relationship between residence times and<br />

either geometrical parameters or hydraulic conditions. Furthermore, Marek et al. (2007)<br />

have po<strong>in</strong>ted out that most of the research <strong>in</strong> flow and mix<strong>in</strong>g processes <strong>in</strong> water storage<br />

tanks has been done with extensive and costly experiments. However, there are few <strong>CFD</strong><br />

studies applied to this sort of modell<strong>in</strong>g <strong>in</strong> the literature (Grayman et al., 1996; Hannoun<br />

and Bolous, 1997; Yeung, 2001; Palau-Salvador et al., 2007).


MATERIAL AND METHODS<br />

The prototype used for the simulations is a 1:4 scale laboratory model of a<br />

rectangular, open-top passage-type water storage tank with constant and equal <strong>in</strong>- and<br />

outflow. Inflow takes place through a circular pipe (<strong>in</strong>ner diameter: 65 mm) as it can be<br />

observed <strong>in</strong> Fig. 1. After a 40° bend, the pipe ends <strong>in</strong> a jet nozzle (nozzle diameter: 16.7<br />

mm). <strong>Water</strong> leaves the tank through a circular pipe (<strong>in</strong>ner diameter: 102 mm), located<br />

close to the <strong>in</strong>let pipe (see Table 1 and Fig. 1). <strong>Water</strong> level is kept constant and no surface<br />

waves were observed dur<strong>in</strong>g the experiments. The experiment beg<strong>in</strong>s with a constant flow<br />

of fresh water (see Table 1). After achiev<strong>in</strong>g an equilibrium with a constant level of the<br />

water surface, br<strong>in</strong>e with a salt concentration of 50 gr/l is <strong>in</strong>jected (t=0) dur<strong>in</strong>g 24 seconds<br />

(the total volume of <strong>in</strong>jected br<strong>in</strong>e is 10 litters).<br />

In this work, three-dimensional simulations are performed on the RANS and<br />

URANS computational techniques <strong>by</strong> us<strong>in</strong>g the STAR-CCM+ commercial code (CD-<br />

Adapco Inc). The numerical results are validated with experimental data which were<br />

carried out at Institute for Hydromechanics, KIT (Karlsruhe). The model is based on the<br />

mass conservation and momentum conservation equations <strong>by</strong> neglect<strong>in</strong>g the temperature<br />

effects. The code is based on a f<strong>in</strong>ite-volume method (see Veersteg and Malalasekera,<br />

1998) for solv<strong>in</strong>g the Navier-Stokes equations. For the purpose of mass conservation a<br />

standard pressure correction algorithm (SIMPLE) is used. For the turbulence closure, a<br />

realizable k-e model was used.<br />

For the <strong>in</strong>let pipe, a velocity <strong>in</strong>let boundary condition was used. For the tank walls<br />

a slip condition was implemented. The free surface was modelled <strong>by</strong> a rigid-lid<br />

approximation. To perform the steady simulations a RANS simulation was performed and<br />

for the unsteady, URANS simulations were carried out. The <strong>in</strong>itial conditions for the<br />

URANS were fixed <strong>by</strong> the steady state of the previous RANS solution. A passive scalar<br />

model was used to simulate the br<strong>in</strong>e <strong>in</strong>jection (STAR-CCM+, 2010). S<strong>in</strong>ce the purpose<br />

of this work was to simulate mix<strong>in</strong>g processes <strong>in</strong>side the tank of a br<strong>in</strong>e <strong>in</strong>jection (with<br />

NaCl as the passive scalar), it was considered a variable-density model.<br />

Different k<strong>in</strong>ds of grids were also developed and tested: structured and<br />

unstructured meshes with different grid sizes. After prelim<strong>in</strong>ary simulations, the best case<br />

among all of them was T1 (structured middle grid). In order to make this selection, a<br />

sensitivity analysis was carried out tak<strong>in</strong>g <strong>in</strong>to account convergence criteria and<br />

simulation time. Therefore, the f<strong>in</strong>al grid used for the calculations has 944,542 grid<br />

po<strong>in</strong>ts. The decrease (stretch<strong>in</strong>g near the wall) is kept to a value around 4%.<br />

RESULTS AND DISCUSSION<br />

Calibrated Model and Validation<br />

Figure 2 shows a comparison between experimental and simulated dilution curves<br />

<strong>in</strong> the outlet cross section of the tank over the first 6,000 seconds. It is possible to observe<br />

that the simulated curve is <strong>in</strong> good agreement with the experimental results. The<br />

beg<strong>in</strong>n<strong>in</strong>g of the first peak of salt concentration, around 80 seconds after the <strong>in</strong>jection, is<br />

fairly good reproduced <strong>by</strong> the simulation <strong>in</strong> time and <strong>in</strong>tensity (the numerical results show<br />

a maximum of 1.04 gr/l versus 1.00 gr/l <strong>in</strong> the experiments). Between the first peak and<br />

until 350 seconds, the model reproduces a s<strong>in</strong>usoidal behaviour. After that, the<br />

concentration decays l<strong>in</strong>early to reach the average value of 0.2 gr/l, mean<strong>in</strong>g an<br />

homogeneous mix<strong>in</strong>g of <strong>in</strong>jection, which is fairly good reproduced <strong>by</strong> the simulation.


Hydrodynamics and <strong>Mix<strong>in</strong>g</strong> <strong>Process</strong><br />

Figure 3 shows the three-dimensional streaml<strong>in</strong>es (coloured <strong>by</strong> the mean velocity<br />

magnitude,<br />

U<br />

<br />

u<br />

2<br />

<br />

v<br />

2<br />

<br />

w<br />

2<br />

) for T1 simulations. These mean flows results<br />

have been obta<strong>in</strong>ed <strong>by</strong> means of RANS simulations. The flow is high three-dimensional<br />

and complex with no dead-zones or clear short-circuit<strong>in</strong>g, lead<strong>in</strong>g to homogeneous<br />

distribution water <strong>in</strong>to the tank. It is possible to observe a significant reduction of velocity<br />

when the jet reaches the opposite corner and how there is clearly a bulk flow circulation<br />

from the entrance to the outlet. It can also be seen two circulations both sides of the jet<br />

and connected above the jet flow.<br />

Figure 4 shows mean velocity contours <strong>in</strong> section Z=0.111. The mean velocity<br />

magnitude is shown <strong>in</strong> each of the grid po<strong>in</strong>ts. Hence, it can be seen clearly the structure<br />

of the flow <strong>in</strong>side the tank and how the <strong>in</strong>let <strong>in</strong>jection <strong>in</strong>fluences the velocity patterns. It is<br />

remarkable the acceleration of the simulated flow, which reaches a maximum value of 2<br />

m/s (location: X=1.64; Y=0.25) compared with the mean <strong>in</strong>let velocity of 0.126 m/s.<br />

Moreover, when the jet strikes on the opposite corner the velocities are reduced and their<br />

values get closer to the mean velocity with<strong>in</strong> the tank (0.0268 m/s). Regard<strong>in</strong>g to the<br />

streaml<strong>in</strong>es, their pattern shows that there are not any short-circuit<strong>in</strong>g as it was expected.<br />

Although the flow is clearly three-dimensional and complex, it is possible to observe at<br />

section Z=0.111 two ma<strong>in</strong> structures both sides of the jet that rotate <strong>in</strong> anticlockwise<br />

direction. It is also possible to identify a relatively small cell near X=1.1 and Y=1.5 that<br />

rotates <strong>in</strong> the opposite direction.<br />

The relevant stagnant zones <strong>in</strong> the tank are plotted <strong>in</strong> Figure 5. The shown isosurface<br />

is determ<strong>in</strong>ed <strong>by</strong> the mean velocity value equal to the 5% of the cumulative<br />

histogram of (0.0056 m/s), represent<strong>in</strong>g then the dead zones with<strong>in</strong> the tank. Two<br />

dead zones stand out clearly: one of them seems to occupy the space between the two<br />

aforementioned ma<strong>in</strong> recirculation structures and the jet; the other one seems to encircle<br />

the jet <strong>in</strong> its right side and may correspond with the core of the second ma<strong>in</strong> recirculation<br />

formed <strong>in</strong> that zone. At the same time, there is another remarkable stagnant zone on the<br />

bottom of the tank and very near the opposite corner of the jet nozzle. Such dead zone has<br />

an elongated shape and may be related to the first ma<strong>in</strong> recirculation (see section Z=0.111<br />

<strong>in</strong> Figure 3). F<strong>in</strong>ally, weaker dead zones can be appreciated, related to typical low<br />

velocity zones <strong>in</strong> tanks such as corners or the free surface, where the transport is limited<br />

<strong>by</strong> the reduced <strong>in</strong>terchange with the mean flow of the tank (Mau et al., 1995).<br />

After the mean velocity results are discussed, salt concentration computed values<br />

<strong>in</strong>side the tank are also showed <strong>in</strong> this section. Figure 6 shows three (1 gr/l, 0.2 gr/l and<br />

0.1 gr/l) iso-surfaces at two different times (13 and 103 seconds s<strong>in</strong>ce the br<strong>in</strong>e <strong>in</strong>jection),<br />

which were supposed to be characteristic of the dilution curve of this specific mix<strong>in</strong>g<br />

process. Therefore, from the snapshots it can be seen that the mix<strong>in</strong>g process is clearly<br />

affected <strong>by</strong> the flow on the dispersion pattern of the scalar <strong>in</strong> the tank, which seems to be<br />

predom<strong>in</strong>antly advective. The 1.00 gr/l iso-surface (at snapshot t=103 seconds) starts<br />

leav<strong>in</strong>g the tank, accord<strong>in</strong>g to the first peak of the dilution curve (see Fig. 2). This means<br />

that an amount of salt concentrations reaches the outlet before it gets completely mixed.<br />

CONCLUSIONS<br />

In this paper, the results of three dimensional numerical simulations have been<br />

presented for a water storage tank. The paper shows a procedure for the analysis of the<br />

mix<strong>in</strong>g processes <strong>in</strong> water tanks us<strong>in</strong>g experimental and computational techniques.


A good agreement has been achieved between experimental and simulated dilution<br />

curve. The mix<strong>in</strong>g process and its relation to the different phases of the dilution curve<br />

have been studied, show<strong>in</strong>g a predom<strong>in</strong>antly advective transport. The relevant stagnant<br />

zones <strong>in</strong> the tank were also located and plotted. The results show a good mix<strong>in</strong>g<br />

capability of the prototype, achiev<strong>in</strong>g a homogeneous mix<strong>in</strong>g of <strong>in</strong>jection after 350<br />

seconds, with no dead-zones or clear short-circuit<strong>in</strong>g <strong>in</strong> the flow.<br />

Future research will focus on the extrapolation ability of the technique, be<strong>in</strong>g<br />

applied, <strong>in</strong> addition with experimental procedures, as a predictive tool on the design,<br />

improvement and management of water tanks.<br />

Literature Cited<br />

Bolous, B. F. et al. 1994. A Discrete Simulation Approach for Network <strong>Water</strong> Quality<br />

Models. J. <strong>Water</strong> Resour. Plng. and Mgmt, ASCE, vol. 121(1), pp. 49–60.<br />

Clark, R. M. et al. 1993. Effect of the Distribution System on the <strong>Water</strong> Quality. J.<br />

AQUA, vol. 42(1), pp. 30–38.<br />

Grayman, W.M. et al. 1996. <strong>Water</strong> Quality and <strong>Mix<strong>in</strong>g</strong> models for tanks and reservoirs.<br />

J.AWWA, 88(7), 60– 73.<br />

Hannoun, I.A. and Boulos, P.F. 1997. Optimiz<strong>in</strong>g Distribution <strong>Storage</strong> <strong>Water</strong> Quality: a<br />

Hydrodynamic Approach. Journal of Applied Mathematical Model<strong>in</strong>g, 21, 495–502.<br />

Marek, M., et al. 2007. <strong>CFD</strong> Model<strong>in</strong>g of Turbulent Jet <strong>Mix<strong>in</strong>g</strong> <strong>in</strong> a <strong>Water</strong> <strong>Storage</strong> <strong>Tank</strong>.<br />

XXXII IAHR Congress, Venice, Italy).<br />

Martínez-Solano, F. J. et al. 2010. Modell<strong>in</strong>g Flow and Concentration Field In rectangular<br />

<strong>Water</strong> <strong>Tank</strong>s. International Congress on Environmental Modell<strong>in</strong>g and Software<br />

Modell<strong>in</strong>g for Environment’s Sake, iEMSs, Fifth Biennial Meet<strong>in</strong>g, Ottawa, Canada.<br />

Mau, R.E. et al. 1995. Explicit Mathematical Models of Distribution <strong>Storage</strong> <strong>Water</strong><br />

Quality. ASCE Journal of Hydraulic Eng<strong>in</strong>eer<strong>in</strong>g, Vol. 121 (10), pp. 669-709.<br />

Palau-Salvador, G. et al. 2007. Numerical <strong>Simulations</strong> to Predict the Hydrodynamics and<br />

the Related <strong>Mix<strong>in</strong>g</strong> <strong>Process</strong>es <strong>in</strong> <strong>Water</strong> <strong>Storage</strong> <strong>Tank</strong>s, Proc. IAHR Congress Venice.<br />

Patwardhan, A.W. 2002. <strong>CFD</strong> Model<strong>in</strong>g of Jet Mixed <strong>Tank</strong>s. Chemical Eng<strong>in</strong>eer<strong>in</strong>g<br />

Science, 57, 1307–1318.<br />

STAR-CCM+ User’s Guide, Version 5.04.006. 2010. CD-Adapco, USA<br />

Veersteg, H.K. and Malalasekera, W. 1998. An Introduction to Computational Fluid<br />

Dynamics. The F<strong>in</strong>ite Volume Method. Longman Scientific and Technical, New<br />

York.<br />

Yeung, H. 2001. Modell<strong>in</strong>g of service reservoirs. Journal of Hydro<strong>in</strong>formatics, 3(3), 165–<br />

172.<br />

Tables<br />

Table 1. Ma<strong>in</strong> Geometrical characteristics and hydraulic conditions of the <strong>Tank</strong>.<br />

Estimated Residence<br />

Inflow<br />

Geometrical Parameters<br />

Time<br />

Q <strong>in</strong><br />

T r<br />

θ φ A B C D E<br />

(l/s)<br />

( h) ( o ) ( o ) (mm) (mm) (mm) (mm) (mm)<br />

0.42 1.58 40 0 150 345 305 150 75<br />

T<br />

<br />

r T<br />

Q <strong>in</strong><br />

, be<strong>in</strong>g <br />

T<br />

<strong>Tank</strong> Volume


Figures<br />

Fig. 1: Detail of the experimental case used for the simulations: T1 case. Left: top view;<br />

right: side view.<br />

Fig. 2: Concentration <strong>in</strong> the outlet section (logarithmic scale) for 6,000 seconds from the<br />

<strong>in</strong>jection of the br<strong>in</strong>e: experimental mean value (dotted curve) and the sectional<br />

averaged value for the simulation (cont<strong>in</strong>uum l<strong>in</strong>e).<br />

Fig. 3: Views of the simulated three-dimensional mean flow field (T1 simulation):<br />

2 2 2<br />

streaml<strong>in</strong>es coloured <strong>by</strong> the mean velocity magnitude ( U u v w ).


Fig. 4: Horizontal sections of mean flow results (T1 simulation). Left: contours of mean<br />

velocity magnitude U<br />

flow <strong>in</strong> section Z=0.111.<br />

2<br />

2<br />

2<br />

u v w <strong>in</strong> section Z=0.111. Right: streaml<strong>in</strong>es of<br />

Fig. 5: Different views of the tank (T1 simulation) show<strong>in</strong>g stagnant zones (5% of the<br />

cumulative histogram of = 0.0056 (m/s)).<br />

Fig. 6: Different views at different specific times (t=13s and t=103s) of the salt<br />

concentration iso-surfaces (1gr/l; 0.2 gr/l and 0.1 gr/l) for T1 simulation.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!